SEMANTICS BOOT CAMP

Transcription

SEMANTICS BOOT CAMP
SEMANTICS
BOOT CAMP
ELIZABETH COPPOCK
Draft of November 27, 2014
Preface
This book is an introduction to formal semantics for beginners,
focussing on the question of how the meaning of a sentence is
built up from the meaning of its part. It synthesizes and distills
the essential points from two classic textbooks in semantics: Introduction to Montague Grammar by Dowty et al. (1981), and Semantics in Generative Grammar by Heim & Kratzer’s (1998). It
also includes some preparatory bits from Mathematical Methods
in Linguistics (Partee et al., 1990). The name comes from a weekend course I taught at the North American Summer School in Language, Logic and Information (NASSLLI) in Austin, Texas in 2012.
The file that produced this book has a direct ancestor in my notes
for that course, although it has come quite a long way since then.
The idea there was to prepare students quickly for the exciting
advanced courses they would be attending at NASSLLI, and the
idea here is similar. This is a comparatively brief introduction that
should allow the student to enter the exciting world of semantics.
The original material comes from my lecture notes and model
solutions from a course called Compositional Semantics which I
taught at Heinrich Heine University in Düsseldorf in 2011 based
on Heim & Kratzer’s textbook, and two semantics courses at the
University of Gothenburg. I thank everyone who participated in
those courses for helping me develop this material. I am also grateful to Robin Cooper for allowing me to sit in on his course on
Montague Semantics at the University of Gothenburg 2012-2013,
which helped me understand it more deeply.
3
4
Heim and Kratzer’s composition rules are all here: Functional
Application, Predicate Modification, Predicate Abstraction, Lexical Terminals, and the Pronouns and Traces Rule. Lambda crunching is involved. The most prominent difference in the framework
is that we translate English expressions into well-formed expressions of lambda calculus rather than specifying denotations directly using an informal metalanguage containing lambdas. People whose minds work like mine may find this more precise style
less puzzling. This style also makes the representation language
more compact, which means that derivations can fit into tree representations. Here is a sample derivation involving both Predicate
Modification and Functional Application:
NP: e
ιx [TEXTBOOK(x ) ∧ ON(SEMANTICS)(x )]
D: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
λP [ιx [P (x )]]
N′ : ⟨e, t ⟩
λx [TEXTBOOK(x ) ∧ ON(SEMANTICS)(x )]
the
N: ⟨e, t ⟩
PP: ⟨e, t ⟩
TEXTBOOK
ON ( SEMANTICS )
textbook
P: ⟨e, ⟨e, t ⟩⟩
NP: e
ON
SEMANTICS
on
semantics
The Predicate Abstraction and Pronouns and Traces rules are
also given a precise formal treatment via translation to lambda
calculus, and it is shown how to deal with quantifier scope ambiguities using Quantifier Raising. However, the empirical argumentation in favor of Quantifier Raising and against a “flexible
5
types” approach that Heim and Kratzer give is presented more
critically.
Another important departure from the Heim & Kratzer framework is in the treatment of presupposition. Partial functions are
replaced with total functions whose range includes an ’undefined’
value, and the partiality operator from Beaver & Krahmer (2001) is
introduced. This means that Functional Application is always defined, and it is easy to read off the presuppositions of a sentence
from its logical translation, and definedness conditions do not get
lost along the way.
There is also a greater emphasis on entailments and the notion of truth relative to a model. The word “model” does not occur
once in the Heim & Kratzer (1998) text, remarkably enough. This
is presumably because the semantics is only meant to be extensional, but leaving models (or possible worlds) out of the picture
leaves us without a method of capturing entailments, and I view
it as the primary job for semantic theory to capture entailments. I
have therefore knit together the model-theoretic approach developed by Richard Montague and laid out in Dowty et al. (1981) with
the more modern and ubiquitous theory of the syntax/semantic
interface laid out by Heim & Kratzer (1998). This style meshes
quite well with the Penn Lambda Calculator (Champollion et al.,
2007), a great pedagogical tool for learning semantics.
While aiming to maintain good formal hygiene, I have also
strived to present things simply and concisely, in order to make
it possible to move to the analysis of linguistic data. Because that
is the fun part.
– Elizabeth Coppock, Uppsala, November 27, 2014
Contents
1 Introduction
11
1.1 Implications . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2 Negative polarity items . . . . . . . . . . . . . . . . . . 15
2 Sets, Relations, and Functions
23
2.1 Set theory . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.1.1 Basics of set theory . . . . . . . . . . . . . . . . 26
2.1.2 Ordered pairs and relations . . . . . . . . . . . 34
3 Functional Application
3.1 Homer snores . . . . . . . .
3.2 Homer loves Maggie . . . .
3.3 Connectives . . . . . . . .
3.4 Tiny Fragment of English .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
39
39
45
50
53
4 Logic
4.1 Propositional calculus . . . . . . . . .
4.1.1 Truth tables . . . . . . . . . . .
4.1.2 Reasoning with truth tables . .
4.1.3 Grammar of PL . . . . . . . . .
4.1.4 Semantics of PL . . . . . . . . .
4.2 First-order logic without variables (L0 )
4.2.1 Syntax of L0 . . . . . . . . . . .
4.2.2 Semantics of L0 . . . . . . . . .
4.3 First-order logic with variables (L1 ) . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
59
59
60
65
68
70
71
72
74
81
7
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
8
CONTENTS
4.3.1 Syntax of L1 . .
4.3.2 Semantics . . .
4.4 Typed lambda calculus
4.4.1 Syntax of L λ . .
4.4.2 Semantics . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
5 Translating to lambda calculus
5.1 Fun with Functional Application . . . . .
5.1.1 Homer is lazy . . . . . . . . . . . .
5.1.2 Homer loves Maggie . . . . . . . . .
5.1.3 Homer is with Maggie . . . . . . . .
5.1.4 Homer is proud of Maggie . . . . .
5.1.5 Homer is a drunkard . . . . . . . .
5.2 Predicate Modification . . . . . . . . . . .
5.2.1 Homer is a lazy drunkard . . . . .
5.3 Quantifiers . . . . . . . . . . . . . . . . . .
5.3.1 Quantifiers: not type e . . . . . . .
5.3.2 Solution: Generalized quantifiers .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
81
85
97
102
103
.
.
.
.
.
.
.
.
.
.
.
109
. 112
. 112
. 114
. 115
. 117
. 118
. 119
. 119
. 122
. 122
. 126
6 Variables
129
6.1 Relative clauses . . . . . . . . . . . . . . . . . . . . . . 129
6.2 Quantifier Raising . . . . . . . . . . . . . . . . . . . . . 133
6.2.1 The problem of quantifiers in object position 134
6.2.2 A Quantifier Raising approach . . . . . . . . . 135
6.2.3 Generalized Argument Raising . . . . . . . . . 138
6.2.4 Putative arguments for the movement . . . . . 141
6.3 Pronouns . . . . . . . . . . . . . . . . . . . . . . . . . . 147
7 Presupposition
7.1 The definite article . . .
7.2 Possessives . . . . . . . .
7.3 Definedness conditions
7.4 Projection problem . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
155
155
160
166
169
CONTENTS
8 Summary
8.1 Logic: Partial typed lambda calculus (L3 )
8.1.1 Syntax of L3 . . . . . . . . . . . . .
8.1.2 Semantics of L3 . . . . . . . . . . .
8.2 Syntax of English fragment . . . . . . . . .
8.3 Translations . . . . . . . . . . . . . . . . .
8.3.1 Lexical entries . . . . . . . . . . . .
8.3.2 Composition rules . . . . . . . . .
9
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
173
174
174
175
177
179
179
182
1
∣
Introduction
1.1 Implications
Semantics and pragmatics are two subfields of linguistics that deal
with meaning. What is meaning? What is it to understand? For instance, does Google understand language? Many might argue that
it does, in some sense. One could point to the fact that one can
type in, “350 USD in SEK” and get back “2232.23 Swedish Krona”
as the first result. This is a step toward understanding. But it can
be argued that in general, Google does not understand language
because it does not do inferences. For example, there is a webpage
that says:
(1)
Beth speaks English, French and Spanish.
If we were to ask Google, “Does Beth speak German?”, Google would
not know the answer. It wouldn’t even know the answer if we
asked, “Does Beth speak a European language?”
A hallmark of a system or agent that understands language or
grasps meaning is that it can do these kinds of inferences. Another way to put this is that a good theory of meaning should
be able to explain when one sentence implies another sentence.
The implication relations that hold between sentences are perhaps the main kind of facts that semantics and pragmatics try to
explain. (Semantics and pragmatics can also be used to explain
why certain sentences sound odd to native speakers, i.e., acceptability judgments, and facts about frequency in natural language
12
Introduction
corpora.)
What exactly does imply mean, i.e., what are implications?1
Implication can be defined either as a relation between two ideas
or as a relation between two sentences. We will define it as a relation between two sentences here. The definition should be weak
enough that it will cover all of the different types of implications
that we will discuss. The following definition will suit our purposes:
(2)
A sentence φ implies another sentence ψ if and only if:
By uttering φ, a speaker signals that ψ is true.
We use φ ‘phi’ and ψ ‘psi’ as placeholders (variables) for sentences,
following the logical tradition of using Greek letters for variables
that range over expressions in the language that is being theorized
about (the object language, which is English here). To apply the
definition, we plug sentences in for these variables. For example:
(1) implies Beth does not speak German, because by uttering Beth
speaks English, French and Spanish, a speaker signals that Beth
does not speak German is true.
There are several different types of implications. Given the
fact that French is a language, (1) logically implies that Beth speaks
a European language. This is an example of an entailment. Entailment is defined as follows:
(3)
A sentence φ entails a sentence ψ if and only if:
Whenever φ is true, ψ is true too.
In the following pairs, the (a) sentence entails the (b) sentence:
(4)
1
a.
All cars are blue.
Terminological aside: An implication (or implication relation) is a relation
that holds between a premise and an implied conclusion. Strictly speaking, the
noun inference describes an act of inferring a conclusion from a premise, but inference can also be used to mean implication. The verb infer is totally different
from the verb imply; a smart person infers a conclusion from a premise, but a
premise implies a conclusion. The subject of infer is the person doing the inference, but the subject of imply is the premise.
Introduction
(5)
(6)
(7)
b.
All sportscars are blue.
a.
Mary invited Fred and Jack.
b.
Mary invited Fred.
a.
I got into MIT yesterday.
b.
I got into MIT.
a.
There are three pens on the table.
b.
More than two pens are on the table.
13
Each of these examples satisfies the definition of entailment. For
example, given that all sportscars are cars, All cars are blue entails
All sportscars are blue, because in every imaginable world or situation where the former is true, the latter is true too. Whenever
There are three pens on the table is true, More than two pens are on
the table is true too.
Note that there are several other ways that entailment can be
defined (Chierchia & McConnell-Ginet, 1990):
(8)
Alternative definitions of entailment from φ to ψ.
a.
The information that ψ conveys is contained in the information that φ conveys.
b.
A situation describable by φ must also be a situation
describable by ψ.
c.
φ and it is not that ψ is contradictory (can’t be true in
any situation).
For example, the information that More than two pens are on the
table conveys is contained in the information that There are three
pens on the table conveys; a situation describable by There are
three pens on the table must also be a situation describable by
More than two pens are on the table, and There are three pens on
the table and it is not true that more than two pens are on the table
cannot be true in any situation.
14
Introduction
Example (1) also implies that Beth does not speak German.
This is not an entailment. It derives from the assumption that the
languages listed make up an exhaustive list of the languages that
Beth speaks. (If she did speak German, the speaker would be “lying by omission”.) This is an example of a conversational implicature. Conversational implicatures are inferences that the hearer
can derive using the assumption that the speaker is adhering to
the norms of conversation.
If I said, Beth no longer speaks French, I would signal a presupposition that she spoke French at some time in the past. To
presuppose something is to take it for granted, to treat it as uncontroversial and known to everyone participating in the conversation. Certain words and phrases, such as no longer, can be used
to signal presuppositions.
Semantics has to do with truth, and how the meaning of a sentence is built up from the meaning of the parts, and thus deals
primarily with entailments. Pragmatics has to do with the interaction between meaning and context. Implicatures are the hallmark topic of pragmatics, and arguably do not fall in the realm of
semantics. There is a great deal of overlap between semantics and
pragmatics, though. The study of presuppositions, for example,
lies in the intersection between semantics and pragmatics. Since
this is a book about semantics, we will focus primarily on entailments, and address presuppositions in the penultimate chapter.
The strategy we will follow to account for entailments is to assign truth conditions to sentences. The idea is that knowing the
meaning of a sentence does not require knowing whether the sentence is in fact true; it only requires being able to discriminate between situations in which the sentence is true and situations in
which the sentence is false. The truth conditions are the conditions under which the sentence is true.
It would not be prudent to claim that meaning of a sentence
in natural language consists entirely in its truth conditions, although this view is sometimes expressed. Heim & Kratzer (1998)
Introduction
15
begin their book with the following bold sentence: “To know the
meaning of a sentence is to know its truth conditions.” Many find
this view objectionable; meaning is not just truth conditions. The
non-entailment implications that a sentence has can also be considered part of its meaning. But truth conditions are a way of capturing one aspect of meaning, namely entailment. If the truth
conditions for ψ are satisfied whenever the truth conditions for
φ are satsified, then we know that φ entails ψ.
How can we assign truth conditions to sentences of natural
languages like English? The philosopher Richard Montague (the
grandfather of modern formal semantics) proposed to treat English as a formal language, like the languages of formal logic. He
famously said, “I reject the contention that an important theoretical difference exists between formal and natural language” (Montague, 1974a, 188). Indeed, there is a very deep connection between logic and semantics. From the beginning, logic has been
oriented toward the question of what makes for a valid argument,
and this of course depends on the notion of entailment. For example, the following is a valid argument:
(9)
Socrates is a man.
All men are mortal.
Therefore, Socrates is mortal.
(Premise 1)
(Premise 2)
(Conclusion)
The first two lines are the premises of the argument and the last
line is the conclusion. The argument is valid because the premises,
taken together, entail the conclusion. In formal semantics, we
make use of tools logicians use to define the concept of validity
and apply them to the problem of accounting for entailment relationships between sentences in natural language.
1.2 Negative polarity items
The concept of entailment is connected to one of the classic puzzles of semantics. Here is the puzzle. There are certain words of
16
Introduction
English, including any, ever, yet, and anymore, which can be used
in negative sentences but not positive sentences:
(10)
Chrysler dealers don’t ever sell any cars anymore.
(11) *Chrysler dealers ever sell any cars anymore.
These are called negative polarity items (NPIs). It’s not just
negative environments where NPIs can be found. Here is a sampling of the data (Ladusaw, 1980).
(12)
⎧
No one
⎪
⎪
⎪
⎪
⎪
At most three people
⎪
⎪
⎪
⎪
⎪ Few students
⎨
*Someone
⎪
⎪
⎪
⎪
⎪
*At least three people
⎪
⎪
⎪
⎪
⎪
⎩ *Many students
⎫
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎬ who had ever read anything about
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎭
phrenology attended any of the lectures.
(13)
⎧
never
⎪
⎪
⎪
⎪
⎪
rarely
⎪
⎪
⎪
⎪
⎪ seldom
I⎨
*usually
⎪
⎪
⎪
⎪
⎪
*always
⎪
⎪
⎪
⎪
⎪
⎩ *sometimes
(14)
a.
John finished his homework {
b.
John voted {
⎫
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎬ ever eat anything for breakfast anymore.
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎭
without
} any help.
*with
against
} ever approving any of the pro*for
posals.
(15)
⎧
before
⎪
⎪
⎪
⎪
⎪ if
John will replace the money ⎨
*after
⎪
⎪
⎪
⎪
⎪
⎩ *when
it.
⎫
⎪
⎪
⎪
⎪
⎪
⎬ anyone ever misses
⎪
⎪
⎪
⎪
⎪
⎭
Introduction
(16)
⎧
hard
⎪
⎪
⎪
⎪
⎪ difficult
It’s ⎨
*easy
⎪
⎪
⎪
⎪
⎪
⎩ *possible
17
⎫
⎪
⎪
⎪
⎪
⎪
⎬ to find anyone who has ever read anything
⎪
⎪
⎪
⎪
⎪
⎭
much about phrenology.
(17)
⎧
doubted
⎪
⎪
⎪
⎪
⎪ denied
John ⎨
*believed
⎪
⎪
⎪
⎪
⎪
⎩ *hoped
⎫
⎪
⎪
⎪
⎪
⎪
⎬ that anyone would ever discover that the
⎪
⎪
⎪
⎪
⎪
⎭
money was missing.
(18)
⎧
is unlikely
⎪
⎪
⎪
⎪
⎪
is doubtful
⎪
⎪
⎪
⎪
⎪
amazed John
⎪
⎪
⎪
It ⎨ *is likely
⎪
⎪
⎪
*is certain
⎪
⎪
⎪
⎪
⎪
is surprising
⎪
⎪
⎪
⎪
⎪
⎩ *is unsurprising
⎫
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎬ that anyone could ever discover that
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎭
the money was missing.
So, some words seem to license negative polarity items and others
do not.
The issue is made a bit more complex by the fact that words
differ as to where they license negative polarity items. No licenses
NPIs throughout the sentence:
(19)
a.
No [ student who had ever read anything about phrenology ] [ attended the lecture ].
b.
No [ student who attended the lectures ] [ had ever read
anything about phrenology ].
But the determiner every licenses NPIs only in the noun phrase
immediately following it:
(20)
a.
Every [ student who had ever read anything about phrenology ] [ attended the lecture ].
18
Introduction
b. *Every [ student who attended the lectures ] [ had ever
read anything about phrenology ].
This shows that the ability to license negative polarity items is not
a simple yes/no matter for each lexical item.
Ladusaw (1980) illustrated a correlation between NPI licensing and “direction of entailment”. A simple, positive sentence containing the word cellist will typically entail the corresponding sentence containing the word musician:
(21)
a.
Mary is a cellist.
b.
⇒ Mary is a musician.
We can describe this situation in terms of sets and subsets. A set
is an abstract collection of distinct objects which are called the
members or elements of that set. Elements may be concrete (like
the beige 1992 Toyota Corolla I sold in 2008, you, or your computer) or abstract (like the number 2, the English phoneme /p/,
or the set of all Swedish soccer players). The elements of a set are
not ordered, and there may be infinitely many of them or none at
all. A set A is a subset of a set B if and only if every member of A
is a member of B .
The set of cellists is a subset of the set of musicians; every cellist is a musician. If we think of these terms as being arranged visually in a taxonomic hierarchy with more specific concepts at the
bottom, we can say that the inference in (5) proceeds from lower
(more specific) to higher (more general), hence “upwards”.
musician
string player
brass player
trombonist
cellist
violinist
...
...
...
Introduction
19
An entailment by a sentence of the form [... A ... ] to a sentence
of the form [ ... B ... ] where A is more specific than B can thus be
labelled an upward entailment. Here is another upward entailment:
(22)
a.
Some cellists snore.
b.
⇒ Some musicians snore.
But negation and the determiner no reverse the entailment pattern:
(23)
(24)
a.
Mary isn’t a musician.
b.
⇒ Mary isn’t a cellist.
a.
No musicians snore.
b.
⇒ No cellists snore.
These entailments are called, as the reader may have guessed, downward entailments, because they go from more general (higher) to
more specific (lower).
There is a correlation between NPI-licensing and downward
entailment: NPIs occur where downward entailments occur. Compare the following examples to the NPI data for no, some and every
above.
(25)
a.
b.
(26)
a.
No musician snores.
⇒ No cellist snores.
(downward)
No musician snores.
⇒ No musician snores loudly.
(downward)
Every musician snores.
⇒ Every cellist snores.
b.
(27)
a.
(downward)
Every musician snores loudly.
⇒ Every musician snores.
(upward)
Some cellists snore.
⇒ Some musicians snore.
(upward)
20
Introduction
b.
Some musician snores loudly.
⇒ Some musician snores.
(upward)
Ladusaw’s generalization was as follows: An expression licenses
negative polarity items in its scope if it licenses downward entailments in its scope. The “scope” of an expression is the constituent
it combines with syntactically. We can assume that a determiner
like no, every, or some combines syntactically with the noun next
to it, and that the resulting noun phrase combines syntactically
with the verb phrase, and the following syntactic structure in particular:2
S
NP
VP
D
N′
no
musician
snores
So no licenses NPIs in the N′ , and the expression no musician licenses NPIs and downward entailments in the VP. Every licenses
NPIs in the N ′ , but the expression every musician does not license
NPIs or downward entailments in the VP.
This generalization works quite well, although there are certain complications. Recall that before and if license NPIs and after
and when do not. So before and if should be downward-entailing
and after and when should not be. Based on the following examples, it seems that after and when are not downward-entailing,
and if is downward-entailing, but before is not, although intuitions
are not entirely clear on this point.
2
S stands for “Sentence”, NP stands for “Noun Phrase”, VP stands for “Verb
Phrase”, D stands for “Determiner”, and N′ should be read “N-bar”; it stands
for an intermediate phrase level in between nouns and noun phrases. Do an
internet search for “X-bar syntax” to find out more about it. The triangles in the
trees indicate that there is additional structure that is not shown in full detail.
Introduction
(28)
(29)
(30)
(31)
21
a.
John will replace the money before he gets to France.
b.
⇒
/ John will replace the money before he gets to Paris.
a.
John will replace the money if he gets to France.
b.
⇒ John will replace the money if he gets to Paris.
a.
John will replace the money after he gets to France.
b.
⇒
/ John will replace the money after he gets to Paris.
a.
John will replace the money when he gets to France.
b.
⇒
/ John will replace the money when he gets to Paris.
Whether or not one gets the intuition that (28a) entails (28b) seems
to depend on whether or not one assumes that John will get to
Paris. Assuming that John will get to Paris, (28a) does seem to
imply (28b), because John will get to France either before he gets
to Paris or simultaneously.3 This is the kind of observation that
motivated von Fintel (1999) to introduce the concept of Strawson
Downward-Entailment (although von Fintel did not address before per se). An environment is Strawson downward-entailing if
it is downward-entailing under the assumption that all of the presuppositions of both sentences are true.
Exercise 1. Check whether Ladusaw’s generalization works for the
other contrasts with respect to NPI licensing that we observed
above – are downward entailments licensed in exactly the places
where NPIs are licensed? (Note that the examples that you need
to construct in order to test this need not contain NPIs; they can
be examples like the ones in (25), (26) and (27).)
3
Similarly, Condoravdi (2010) observes that Ed left before we were in the room
does not intuitively imply Ed left before we were in the room standing by the window, but the inference does go through if we assume that we were standing by
the window.
2
∣
Sets, Relations, and Functions
Recall Ladusaw’s generalization regarding negative polarity item
licensing: An expression licenses negative polarity items in its scope
if it licenses downward entailments in its scope, where the “scope”
of an expression is the constituent it combines with syntactically,
its syntactic “sister”. Ladusaw also gives an explanation why the
determiners some, no and every license the upward and downward entailments that they do. It starts with the idea that these
words can be thought of as relations between two sets.
• every can be thought of as a relation between two sets X and
Y which holds if X is a subset of Y , i.e., if every member of X
is a member of Y . The sentence every musician snores, for
instance, expresses that every member of the set of musicians is a member of the set of people who snore. This type
of scenario can be depicted as follows, with the blue circle
for the snorers and the red circle for the musicians.
• some can be thought of as a relation between two sets X and
Y which holds if there is some member of X which is also a
member of Y , i.e., if the intersection between X and Y is
non-empty. For instance, some musician snores should be
24
Sets, Relations, and Functions
true if there is some individual which is both a musician and
a snorer. This type of scenario looks like this:
• no can be thought of as a relation between two sets X and
Y which holds if the two sets have no members in common,
in other words, if the intersection is empty. So no musician
snores holds if there is no individual who is both a musician
and a snorer. In that case, the two sets are disjoint, like so:
Consider what happens when we consider a subset X ′ of X
(e.g., the set of cellists). Every X Y means that X is a subset of Y .
If that is true, then any subset X ′ of X will also be a subset of Y .
This can be visualized as follows, with the innermost circle representing the cellists.
So, for example, if Every musician snores is true, then Every cellist
snores is also true. Since every X Y entails every X ′ Y for every X ′
that is a subset of X , we can say that every is left downward monotone (“left” because it has to do with the element on the left, X ,
Sets, Relations, and Functions
25
rather than the element on the right, Y .) In general, a determiner
δ is left downward monotone if δX Y entails δX ′ Y for all X ′ that
are subsets of X .
A determiner δ is right downward monotone if δX Y entails
δX Y ′ for any Y ′ that is a subset of Y . Let us consider whether
every is right downward monotone. Suppose that every X Y is true.
Then X is a subset of Y . Now we will take a subset of Y , Y ′ . Are we
guaranteed that X is a subset of Y ′ ? No! Consider the following
scenario.
Or think about it this way: Just because every musician snores
doesn’t mean that every musician snores loudly. So every is not
right downward monotone.
Now let us consider some. With some, we are not guaranteed
that the sentence will remain true when we replace X with a subset X ′ . Some X Y means that the intersection of X and Y contains
at least one member. If we take a subset X ′ of X , then we might
end up with a set that has no members in common with Y , like
this:
So, for example, suppose that Some musician snores is true. This
does not mean that Some cellist snores is true, because it could
be the case that none of the musicians who snore are cellists. So
some is not left downward monotone. By analogous reasoning, it
isn’t right downward monotone either.
26
Sets, Relations, and Functions
Exercise 1. Is no left downward monotone? Is it right downward
monotone? Explain.
2.1 Set theory
2.1.1 Basics of set theory
In the previous section, we used several concepts, including set,
subset, and intersection. Here is a fuller listing of the concepts of
set theory that play an important role in semantics, along with the
notation that we typically use for them.
A set is an abstract collection of distinct objects which are called
the members or elements of that set. Here is a passage from Partee et al. (1990) introducing the notion of a set:
Objects of quite different nature can be members of a
set, e.g. the set of red objects may contain cars, bloodcells, or painted representations. Members of a set
may be concrete, like cars, blood-cells, or physical sounds,
or they may be abstractions of some sort, like the number two, or the English phoneme /p/, or a sentence
of Chinese. In fact, we may arbitrarily collect objects
into a set even though they share no property other
than being a member of that set...
Sets may be large, e.g. the set of human beings, or
small, e.g. the set of authors of this book. Sets are either finite, e.g., the readers of this book or the set of
natural numbers between 2 and 98407, or they are infinite, e.g. the set of sentences of a natural language
or the set of natural numbers: zero, one, two, three, ...
Since members of sets may be abstract objects, a set
may in particular have another set as a member. A set
can thus simultaneously be a member of another set
Sets, Relations, and Functions
27
and have other sets as members. This characteristic
makes set theory a very powerful tool for mathematical and linguistic analysis.
If a set has only one member, it is called a singleton. For example,
the set containing only the number two is a singleton. The set with
no elements at all is called the empty set, written ∅.
Partee et al. (1990) also point out:
A set may be a legitimate object even when our knowledge of its membership is uncertain or incomplete.
The set of Roman emperors is well-defined even though
its membership is not widely known, and similarly the
set of all former first-grade teachers is perfectly determined, although it may be hard to find out who belongs to it. For a set to be well-defined it must be clear
in principle what makes an object qualify as a member of it...
If you know the identities of the elements of the set and have the
time and space to list them out, then you can specify a set using
list notation, with curly braces. For example:
{2, 4, 6}
denotes the set containing 2, 4 and 6. One can also use list notation to denote the empty set, simply by not listing any members:
{} is another way of referring to the empty set.
If you wish to refer to the set of objects satisfying a certain condition, then you can use predicate notation. For example:
{x ∣ x is a positive even number less than 7}
denotes the set of positive even numbers less than 7 (i.e., the set
containing 2, 4 and 6, the same set as in the previous example).
The vertical bar means “such that”, so the formula can be read,
“the set of x such that x is a positive even number less than 7”.
28
Sets, Relations, and Functions
To express the idea that some entity x is an element of set A,
we use the special symbol ∈ thus:
x∈A
This means ‘x is an element of A’. (We typically use capital letters like A, B , C , etc. as names for arbitrary sets that we posit for
the sake of discussion. Members of sets typically get names with
lower-case letters.) We can also write:
x ∈/ A
for ‘x is not an element of A’. For example, the following are true
statements:
2 ∈ {2, 4, 6}
3 ∈/ {2, 4, 6}
(assuming that ‘2’ picks out the number two, etc.).
There are a few other symbols that are commonly used to talk
about sets. To express that A is a subset of B , we write A ⊆ B . The
official definition of subset is as follows: A is a subset of B if and
only if every member of A is a member of B . In symbols:
A ⊆ B iff for all x: if x ∈ A then x ∈ B .
For example, the following statements are true:
{a, b } ⊆ {a, b, c }
{b, c } ⊆ {a, b, c }
{a } ⊆ {a, b, c }
∅ ⊆ {a, b, c }
{a, b, c } ⊆ {a, b, c }
Things get slightly trickier to think about when the elements of the
sets involved are themselves sets. Here is another true statement:
{a, {b }} ⊆/ {a, b, c }
Sets, Relations, and Functions
29
The reason {a, {b }} is not a subset of {a, b, c } is that the former
has a member that is not a member of the latter, namely {b }. It is
tempting to think that {a, {b }} contains b but this is not correct.
The set {a, {b }} has exactly two elements, namely: a and {b }. The
set {b } is not the same thing as b. One is a set and the other might
not be. The following is a true statement, though:
{a, {b }} ⊆ {a, {b }, c }
Every element of {a, {b }} is an element of {a, {b }, c }, as we can
see by observing that the following two statements hold:
a ∈ {a, {b }, c }
{b } ∈ {a, {b }, c }
Note that the empty set is a subset (not an element!) of every set.
Since the empty set doesn’t have any members, it never contains
anything that is not part of another set, so the definition of subset
is always trivially satisfied. So whenever anybody asks you, “Is the
empty set a subset of...?”, you can answer “yes” without even hearing the rest of the sentence. (If they ask you whether the empty set
is an element of some other set, then you’ll have to look among the
elements of the set in order to decide.)
Note also that by this definition, every set is actually a subset
of itself, even though normally we think of two sets of different
sizes when we think of the subset relation. To avoid confusion, it
helps to distinguish between subsets and proper subsets. A is a
proper subset of B , written A ⊂ B , if and only if A is a subset of B
and A is not equal to B .
A ⊂ B iff (i) for all x: if x ∈ A then x ∈ B and (ii) A ≠ B .
For example, {a, b, c } ⊆ {a, b, c } but it is not the case that {a, b, c } ⊂
{a, b, c }.
The reverse of subset is superset. A is a superset of B , written
A ⊇ B , if and only if every member of B is ia member of A.
30
Sets, Relations, and Functions
A ⊇ B iff for all x: if x ∈ B then x ∈ A.
And as you might expect, A is a proper superset of B , written A ⊃ B ,
if and only if A is a superset of B and A is not equal to B .
A ⊃ B iff (i) for all x: if x ∈ B then x ∈ A and (ii) A ≠ B .
The intersection of A and B , written A ∩ B , is the set of all entities x such that x is a member of A and x is a member of B .
A ∩ B = {x ∣ x ∈ A and x ∈ B }
For example:
{a, b, c } ∩ {b, c, d } = {b, c }
{b } ∩ {b, c, d } = {b }
{a } ∩ {b, c, d } = ∅
{a, b } ∩ {a, b } = {a, b }
Intersection is very useful in natural language semantics. It can be
used as the basis for a semantics of and. For example, if someone
tells you that John is a lawyer and a doctor, then you know that
John is in the intersection between the set of lawyers and the set of
doctors. So John is in the purple area in the Venn diagram, where
doctors and lawyers overlap.
doctors
lawyers
Sets, Relations, and Functions
31
Another useful operation on sets is union. The union of A and
B , written A ∪ B , is the set of all entities x such that x is a member
of A or x is a member of B .
A ∪ B = {x ∣ x ∈ A or x ∈ B }
For example:
{a, b } ∪ {d , e } = {a, b, d , e }
{a, b } ∪ {b, c } = {a, b, c }
{a, b } ∪ ∅ = {a, b }
As the reader can guess, union can be used to give a semantics for
or. If someone tells you that John is a lawyer or a doctor, then you
know that John is in the union of the set of lawyers and the set of
doctors. This would place him somewhere in one of the shaded
areas in the Venn diagram, be it red (doctor, but not lawyer), blue
(lawyer, but not doctor), or purple (both). (You might normally
assume that he is not in the intersection of doctors and lawyers
though – that he is either a doctor or a lawyer, but not both. This
is called an exclusive interpretation for or, and we will get to that
later on.)
Exercises on sets
The following exercises are taken from Partee, ter Meulen and Wall,
Mathematical Methods in Linguistics.
Exercise 2. Given the following sets:
A = {a, b, c, 2, 3, 4}
B = {a, b }
C = {c, 2}
D = {b, c }
E = {a, b, {c }}
F =∅
G = {{a, b }, {c, 2}}
classify each of the following statements as true or false.
32
Sets, Relations, and Functions
(a) c ∈ A
(b) c ∈ F
(c) c ∈ E
(d) {c } ∈ E
(e) {c } ∈ C
(f) B ⊆ A
(g) D ⊂ A
(h) A ⊆ C
(i) D ⊆ E
(j) F ⊆ A
(k) E ⊆ F
(l) B ∈ G
(m) B ⊆ G
(n) {B } ⊆ G
(o) D ⊆ G
(p) {D } ⊆ G
(q) G ⊆ A
(r) {{c }} ⊆ E
Exercise 3. Consider the following sets:
S 1 = {{∅}, { A }, A }
S2 = A
S3 = { A}
S 4 = {{ A }}
S 5 = {{ A }, A }
S6 = ∅
S 7 = {∅}
S 8 = {{∅}}
S 9 = {∅, {∅}}
(a) Of the sets S 1 − S 9 , which are members of S 1 ?
(b) Which are subsets of S 1 ?
(c) Which are members of S 9 ?
(d) Which are subsets of S 9 ?
(e) Which are members of S 4 ?
(f) Which are subsets of S 4 ?
Exercise 4. Given the sets A, ...,G from above, repeated here:
A = {a, b, c, 2, 3, 4}
B = {a, b }
C = {c, 2}
D = {b, c }
E = {a, b, {c }}
F =∅
G = {{a, b }, {c, 2}}
Sets, Relations, and Functions
33
list the members of each of the following:
(a) B ∪ C
(g) A ∩ E
(m) B − A
(b) A ∪ B
(h) C ∩ D
(n) C − D
(c) D ∪ E
(i) B ∩ F
(o) E − F
(d) B ∪ G
(j) C ∩ E
(p) F − A
(e) D ∪ F
(k) B ∩ G
(f) A ∩ B
(l) A − B
(q) G − B
Exercise 5. Let A = {a, b, c }, B = {c, d }, C = {d , e, f }. Calculate the
following:
(a) A ∪ B
(b) A ∩ B
(c) A ∪ (B ∩ C )
(d) C ∪ A
(e) B ∪ ∅
(f) A ∩ (B ∩ C )
(g) A − B
(h) Is a a member of { A, B }?
(i) Is a a member of A ∪ B ?
34
Sets, Relations, and Functions
2.1.2 Ordered pairs and relations
The meanings of common nouns like cellist and intransitive verbs
like snores are often thought of as sets (the set of cellists, the set
of individuals who snore, etc.). Transitive verbs like love, admire,
and respect are sometimes thought of as relations between two individuals. Technically, a relation is a set of ordered pairs.
Ordered pair.
As stated above, sets are not ordered:
{a, b } = {b, a }
But the elements of an ordered pair written ⟨a, b ⟩ are ordered.
Here, a is the first member and b is the second member.
⟨a, b ⟩ ≠ ⟨b, a ⟩
The notion of ordered pair can be defined in terms of the notion
of set. We can define ⟨a, b ⟩ as {a, {a, b }}. Alternatively, we can
think of ordered pairs as a separate primitive.
Exercise 6. True or false?
(a) {3, 3} = {3}
(b) {3, 4} = {4, 3}
(c) ⟨3, 4⟩ = ⟨4, 3⟩
(d) ⟨3, 3⟩ = ⟨3, 3⟩
(e) ⟨3, 3⟩ = ⟨3⟩
(f) {⟨3, 3⟩} = ⟨3, 3⟩
(g) {⟨3, 3⟩, ⟨3, 4⟩} = {⟨3, 4⟩, ⟨3, 3⟩}
(h) ⟨3, {3, 4}⟩ = ⟨3, {4, 3}⟩
(i) {3, {3, 4}} = {3, {4, 3}}
Sets, Relations, and Functions
35
Cartesian product. The Cartesian product of two sets A and B ,
written A × B , is the set of ordered pairs that can be constructed
by pairing a member of A with a member of B .
A × B = {⟨x, y ⟩ ∣ x ∈ A and y ∈ B}
For example, the Cartesian product of {a, b, c } and {1, 0} is:
{⟨a, 1⟩, ⟨a, 0⟩, ⟨b, 1⟩, ⟨b, 0⟩, ⟨c, 1⟩, ⟨c, 0⟩}
Exercise 7. Let D e be the set containing the members of the Simpsons family:
D e = {Homer, Marge, Bart, Lisa, Maggie}
Let D t be {0, 1}. The idea behind this notation is that e represents
the type of individual entities, and t represents the type of truth
values (0 for false, 1 for true). D stands for “domain”.
(a) What is the Cartesian product of D e and D t ?
(b) What is the Cartesian product of D t and itself?
Relation. As mentioned above, transitive verbs like love, admire,
and respect are sometimes thought of as relations between two individuals. Certain nouns like mother can also be thought of as
expressing relations. The love relation can be treated as the set of
ordered pairs of individuals such that the first member loves the
second member. Suppose John loves Mary. Then the pair ⟨John,
Mary⟩ is a member of the love relation.
A relation has a domain and a range. The domain is the set
of objects from which the first members are drawn, and the range
is the set of objects from which the second members are drawn.
If R is a relation between two sets A and B , then we say that R is
36
Sets, Relations, and Functions
a relation from A to B . If R is a relation from A to B , then R is a
subset of the Cartesian product of A and B , i.e. R ⊆ A × B .
Exercise 8. Consider the following relations from the set containing the Nordic countries (Iceland, Sweden, Norway, Finland, and
Denmark) to {0, 1}:
R 1 : {⟨Sweden,0⟩, ⟨Norway,0⟩, ⟨Iceland,1⟩, ⟨Demmark,1⟩, ⟨Finland,1⟩}
R 2 ∶ {⟨Sweden,1⟩, ⟨Norway,1⟩, ⟨Iceland,0⟩, ⟨Demmark,0⟩, ⟨Finland,0⟩}
R 3 : {⟨Sweden,1⟩, ⟨Norway,1⟩}
R 4 ∶ {⟨Sweden,1⟩, ⟨Norway,1⟩, ⟨Iceland,0⟩, ⟨Demmark,0⟩, ⟨Finland,0⟩,
⟨Sweden,0⟩, ⟨Norway,0⟩}
R 5 : {Sweden, Norway, Iceland, Denmark, Finland} × {0, 1}
Questions:
(a) All of these relations have the same domain. What is it?
(b) All of these relations have the same range. What is it?
(c) True or false:
All of these relations are subsets of the Cartesian product of
{Iceland, Sweden, Norway, Finland, Denmark} and {0, 1}.
Function. A function is a special kind of relation. A relation R
from A to B is a function if and only if it meets both of the following conditions:
• Each element in the domain is paired with just one element
in the range.
Sets, Relations, and Functions
37
• The domain of R is equal to A.
A function can be thought of as something that gives a single output for a given input. For example, {⟨0, 0⟩, ⟨0, 1⟩} is not a function
because given the input 0, there are two outputs: 0 and 1.
Exercise 9. Which of the relations R 1 − R 5 are functions?
We write F (a ) to denote ‘the result of applying function F to
argument a’ or F of a’ or ‘F applied to a’. If F is a function that
contains the ordered pair ⟨a, b ⟩, then:
F (a ) = b
This means that given a as input, F gives b as output.
Exercise 10. Given R 1 as defined above:
(a) What value does R 1 take, given Norway as an argument?
(b) In other words, what is R 1 (Norway)?
(c) What is R 1 (Iceland)?
(d) What is R 2 (Norway)?
Characteristic function. As mentioned above, the semantics of
common nouns like tiger and picnic and student are sometimes
thought of in terms of sets – the set of tigers, the set of picnics, the
set of students. But another way of treating their meaning is as
functions from individuals to truth values. For example, the set of
children in the Simpsons family is {Bart, Maggie, Lisa}. One possible analysis of the word child relative to the Simpsons domain
38
Sets, Relations, and Functions
is this set. But we could also treat it as a function that takes an
individual and returns a truth value (0 or 1; 0 for “false” and 1 for
“true”):
{⟨Homer, 0⟩, ⟨Marge, 0⟩, ⟨Bart, 1⟩, ⟨Lisa, 1⟩, ⟨Maggie, 1⟩}
Here is an alternative way of writing the same thing:
⎡
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎣
Homer
Marge
Bart
Lisa
Maggie
→
→
→
→
→
0
0
1
1
1
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
This function, applied to Lisa, would yield 1 (“true”). Applied to
Homer, it would yield 0 (“false”), because Homer is not a child (despite the fact that Lisa is more mature than he is). A function that
yields 1 (“true”) for every input that is in set S is called the characteristic function of S. The function we have just been looking at
is the characteristic function of the set {Bart, Lisa, Maggie}.
Exercise 11.
(a) What set is R 1 is the characteristic function of? (You can define
it by listing the elements.)
(b) What set is R 2 is the characteristic function of?
(c) Give the characteristic function of the set of Scandinavian
countries, with the set of Nordic countries as its domain.
3
∣
Functional Application
3.1 Homer snores
Now we will begin to talk about natural language expressions and
their denotations. We use denotation brackets ⟦⋅⟧ to connect natural language expressions with their denotations. The denotation
brackets represent a function that maps natural language expressions the individuals, sets, relations, etc. that they denote. Suppose that the denotation of the noun phrase Homer is the individual Homer Simpson. We can write this as follows:
⟦Homer⟧ = Homer Simpson
Note that we use italics when we are talking about English. The
language that we are talking about is called the object language
(as opposed to the meta-language, which is the language we use
to talk about the object language). Throughout this book, italics
are used for object language (as well as emphasis). Suppose that
the denotation of the verb snores is as follows:
⟦snores⟧ = {⟨Homer, 1⟩, ⟨Marge, 1⟩, ⟨Bart, 1⟩, ⟨Lisa, 0⟩, ⟨Maggie, 0⟩}
Here is another way of writing the same thing:
⎡
⎢
⎢
⎢
⎢
⟦snores⟧ = ⎢
⎢
⎢
⎢
⎢
⎣
Homer
Marge
Bart
Lisa
Maggie
39
→
→
→
→
→
1
1
1
0
0
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
40
Functional Application
This function, when applied to Homer, equals 1:
⎡
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎣
Homer
Marge
Bart
Lisa
Maggie
→
→
→
→
→
1
1
1
0
0
⎤
⎥
⎥
⎥
⎥
⎥ (Homer) = 1
⎥
⎥
⎥
⎥
⎦
Hence ⟦snores⟧(Homer)= 1, which in turn means that:
⟦snores⟧(⟦Homer⟧)= 1.
Exercise 1. Assume that ⟦Marge⟧ = Marge.
(a) What is the value of ⟦snores⟧(⟦Marge⟧)?
(b) What is the value of ⟦snores⟧(Marge)?
(c) Does it make any sense to write ⟦snores⟧ (Marge)? Why or why
not?
Now let us consider how to define the denotation of a sentence like Homer snores. Rather than assigning denotations to sequences of words, we will assign denotations to structural analyses of sentences (syntax trees), in order to avoid having to deal
with syntactic ambiguity when we are doing semantics. Let us begin with a simple grammar that generates sentences like Homer
snores and Homer loves Marge. It says that a sentence (S) can be
made up of a noun phrase (NP) and a verb phrase (VP), and a verb
phrase can be made up of a verb (V) and a noun phrase (NP), or
just a verb. (The parentheses indicate optionality.)
(1)
Syntax
S
→
VP →
NP VP
V (NP)
Functional Application
41
Here we are using rewrite rules in the style of context-free grammar, a practice introduced into linguistics in the 1950s by Noam
Chomsky (see for example Chomsky 1957, 1965). The expressions
of category S are the sentences of the language. We may assume
in addition the following simple lexicon (specifying which words
of which categories there are).
(2)
Lexicon
NP: Bart, Maggie, Homer, Marge, Lisa
V: snores, loves
This grammar generates analysis trees including the following:
S
S
NP
VP
NP
Homer
V
Homer
snores
VP
V
NP
loves
Maggie
It is trees like this that we will place inside denotation brackets,
like so:
(3)
L
M
S
P
Q
P
Q
P
Q
P NP
VP Q
P
Q
P
Q
P
Q
P Homer
Q
V
P
Q
P
Q
P
Q
P
Q
snores
N
O
(So technically, this will be undefined: ⟦Homer snores⟧.)
We said above that ⟦Homer⟧ is the individual Homer Simpson,
and ⟦snores⟧ is a function which, when applied to Homer, yields
the value 1 (“true”). So intuitively, the sentence is true. This means
that the value of (3) should be the result of applying ⟦snores⟧ to
⟦Homer⟧. This is the result that should come out, intuitively. But
so far, we have not specified any rules in our semantic system that
42
Functional Application
would derive this result. We can fill this gap by adopting the principle of Functional Application:
Composition Rule 1: Functional Application (FA). If γ is a tree
whose only two subtrees are α and β, and ⟦β⟧ is a function whose
domain contains ⟦γ⟧, then ⟦α⟧= ⟦β⟧(⟦γ⟧).
Notice that we are using Greek letters (α ‘alpha’, β ‘beta’, and γ
‘gamma’) as variables that range over expressions of the object
language (trees). The tree inside the denotation brackets in (3)
has two subtrees, the NP and the VP. The VP denotes a function
whose domain includes the denotation of the NP. Hence the rule
of Functional Application implies that the denotation of the S-tree
is the denotation of the VP-tree applied to the denotation of the
NP-tree:
L
S
P
P
P
P NP
VP
P
P
P
P Homer
V
P
P
P
P
snores
N
M
Q
Q L VP
Q P
Q P
Q P
Q=P V
Q P
Q P
Q P
Q P snores
Q N
Q
O
M
Q
Q L NP M
Q ⎛P
Q⎞
Q P
Q
Q ⎝P
Q
Q P Homer Q⎠
Q N
O
Q
O
The parentheses in the above expression signify the result of applying a function to its argument.
We still haven’t said enough to compute the meaning of the
sentence. We know what ⟦Homer⟧ is but we haven’t said yet what
L NP M
P
Q
P
Q
P
Q
P Homer Q
N
O
is. To remedy this, let us add a rule for unary branching nodes
(nodes in the tree that have only one daughter), letting the semantic value be passed up from daughter to mother in case there is
only one daughter:
Functional Application
43
Composition Rule 2: Non-branching Nodes (NN). If α is a tree
whose only subtree is β, then ⟦α⟧=⟦β⟧.
The Non-branching Nodes rule tells us that:
L NP
P
P
P
P Homer
N
M
Q
Q = ⟦Homer⟧
Q
Q
O
and that:
L VP
P
P
P
P V
P
P
P
P snores
N
M
Q
Q L V M
Q P
Q
Q=P
Q = ⟦snores⟧
Q P
Q
Q P snores Q
Q N
O
Q
O
Now we are in a position to calculate the denotation for the
whole sentence:
44
Functional Application
L
M
S
P
Q
P
Q
P
Q
P NP
Q
VP
P
Q
P
Q
P
Q
P Homer
Q
V
P
Q
P
Q
P
Q
P
snores Q
N
O
L VP M
P
Q
P
Q L NP M
P
Q ⎛P
Q⎞
P
P
Q
= P V Q
Q ⎝P
Q
P
Q P Homer Q⎠
P
Q N
O
P snores Q
N
O
= ⟦snores⟧(⟦Homer⟧)
⎡
⎢
⎢
⎢
⎢
= ⎢
⎢
⎢
⎢
⎢
⎣
= 1
Homer
Marge
Bart
Lisa
Maggie
→
→
→
→
→
1
1
1
0
0
⎤
⎥
⎥
⎥
⎥
⎥ (Homer)
⎥
⎥
⎥
⎥
⎦
Note that we used the Non-branching Nodes rule several times at
once in the transition between the second line and the third line.
Exercise 2. How many times was Non-branching Nodes used in
the transition between the second line and the third line?
Exercise 3. Do a step-by-step calculation of the semantic value
of Bart snores, assuming that ⟦Bart⟧=Bart, and show your work.
Mention the composition rules (Functional Application, Nonbranching Nodes) that you use at each step.
Functional Application
45
3.2 Homer loves Maggie
Let us assume that loves Maggie also denotes the characteristic
function of a set, the set of individuals who love Maggie. Then
Homer will be able to combine with loves Maggie in the same way
that it combines with snores. Let us assume that every individual
loves Maggie except for Bart. So loves Maggie should turn out to
denote the following function:
(4)
{⟨Homer, 1⟩, ⟨Marge, 1⟩, ⟨Bart, 0⟩, ⟨Lisa, 1⟩, ⟨Maggie, 1⟩}
Alternative notation:
⎡
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎣
Homer
Marge
Bart
Lisa
Maggie
→
→
→
→
→
1
1
0
1
1
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
For example, since Homer loves Maggie, the denotation of loves
Maggie applied to Homer equals 1 (true).
What meaning can we assign to loves so that loves Maggie has
(4) as its denotation? If we analyze a transitive verb like loves as a
function, then we can use Functional Application to put together
the meaning of a transitive verb with its object as well. The result of applying that function to its direct object should itself be
a function that can combine via Functional Application with the
subject. So loves should denote a function which, when applied
to the denotation of Maggie, yields (4). So it should turn out to be
the case that:
(5)
⟦loves⟧(⟦Maggie⟧)
= {⟨Homer, 1⟩, ⟨Marge, 1⟩, ⟨Bart, 0⟩, ⟨Lisa, 1⟩, ⟨Maggie, 1⟩}
At this point, it is useful to introduce the notion of semantic
type. This gives us a vocabulary for distinguishing among the different kinds of denotations that expressions of natural language
may have. Let us distinguish two basic types:
46
Functional Application
• Individuals (like Bart and Lisa) are type e (for “entity”);
• Truth values (0 and 1) are type t .
We refer to the set of individuals as D e ‘the domain of individuals’ and the set of truth values as D t ‘the domain of truth values’. Functions also have types. A function whose domain is D e
and whose range is D t is of type ⟨e, t ⟩. For example, both snores
and loves Maggie denote functions of type ⟨e, t ⟩. A function from
truth-values to truth-values is of type ⟨t , t ⟩. Here is an example of
a function of type ⟨t , t ⟩:
{⟨1, 0⟩, ⟨0, 0⟩}
This function takes an input of type t and gives an output of type
t . Here is a function of type ⟨t , ⟨t , t ⟩⟩:
{⟨0, {⟨0, 0⟩, ⟨1, 0⟩}⟩, ⟨1, {⟨0, 0⟩, ⟨1, 1⟩}⟩}
It takes an input of type t and gives an output of type ⟨t , t ⟩. For
example, given the input 0, it returns {⟨0, 0⟩, ⟨1, 0⟩}, a function of
type ⟨t , t ⟩. Since the input is of type t and the output is of type
⟨t , t ⟩, the type of the function is ⟨t , ⟨t , t ⟩⟩.
In general, a function whose domain is D a for some type a and
whose range is D b for some type b is of type ⟨a, b ⟩. In other words,
for any types a and b, a function from D a to D b is of type ⟨a, b ⟩.
For any types a and b, let us say that a function of type ⟨a, b ⟩ is
in D ⟨a,b ⟩ . Our inventory of types can be defined recursively1 as
follows:
• e is a type.
• t is a type.
1
A recursive definition uses the term to be defined in the definition, but it is
not circular because there is a so-called base case. Readers may be familiar with
the ‘factorial’ function, e.g. 5! = 5 × 4 × 3 × 2 × 1. This can be defined recursively:
If n = 0, then n! = 1 (base case). Otherwise, n! = n × (n − 1)! (recursive case).
Functional Application
47
• For any types a and b, ⟨a, b ⟩ is a type.
• Nothing else is a type.
Notice that by these definitions, angle brackets only appear with
complex types of the form ⟨a, b ⟩, not the simple types e and t .
Exercise 4.
(a) What type is Bart?
(b) What type is 0?
(c) What type is this function: {⟨0, 0⟩, ⟨1, 0⟩}
(Assume that the range of the function is D t .)
(d) Give an example of a function of type ⟨t , ⟨t , t ⟩⟩.
(e) Is ⟨⟨e, t ⟩, t ⟩ a type?
(f) Is ⟨e, ⟨e, t ⟩⟩ a type?
(g) If a function is of type ⟨e, ⟨e, t ⟩⟩, what is its domain? What is
its range?
(h) Is ⟨⟨e, t ⟩, ⟨t , ⟨t , t ⟩⟩⟩ a type?
(i) Is ⟨⟨e, t ⟩⟩ a type?
(j) Is ⟨e ⟩ a type?
(k) Is e a type?
Let us return to the main thread. The verb loves should denote a function which, in general, given an individual of type e as
an argument, returns a function of type ⟨e, t ⟩ as a result. Suppose
that the state of the world is as follows. Everybody but Bart loves
48
Functional Application
Maggie, everybody loves themselves, and nobody else loves anybody else. Then the denotation of loves, ⟦loves⟧, is as in Figure 3.1:
The first argument of this function represents the person who is
loved, and the second argument represents the lover. This works
out nicely with the fact that the verb forms a syntactic constituent
with the direct object, the position where the person who is loved
is represented. Functional Application will dictate that the verb
combines first with the direct object, and the verb phrase then
combines with the subject. Here is a schematic representation of
how the derivation for a sentence like Homer loves Maggie will go:
t
(6)
e
Homer
⟨e, t ⟩
⟨e, ⟨e, t ⟩⟩
e
loves
Maggie
The transitive verb denotes a function that takes an individual
(Maggie in this case), and returns a function that takes an individual (Homer in this case), and returns a truth value (0 or 1). The
input to the denotation of the transitive verb is an individual, type
e. The output is the characteristic function of a set of individuals,
type ⟨e, t ⟩. Since the input is type e, and the output is type ⟨e, t ⟩,
the function is type ⟨e, ⟨e, t ⟩⟩. This is why we have ⟨e, ⟨e, t ⟩⟩ above
loves in (6). When this function applies to Maggie, the result will
be a function of type ⟨e, t ⟩. That’s why the type for the VP loves
Maggie is ⟨e, t ⟩. This is the perfect type to combine via Functional
Application with the subject. Just like an intransitive verb, loves
Maggie combines with an individual (type e) to produce a truth
value (type t ).
Notice that we have used Functional Application in two different syntactic environments. We used it to put a subject NP together with a VP, and in that case the constituent on the right was
Functional Application
⎡
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎣
⎡
⎢
⎢
⎢
⎢
Maggie → ⎢
⎢
⎢
⎢
⎢
⎣
⎡
⎢
⎢
⎢
⎢
Lisa
→⎢
⎢
⎢
⎢
⎢
⎣
⎡
⎢
⎢
⎢
⎢
Bart
→⎢
⎢
⎢
⎢
⎢
⎣
⎡
⎢
⎢
⎢
⎢
Homer → ⎢
⎢
⎢
⎢
⎢
⎣
⎡
⎢
⎢
⎢
⎢
Homer → ⎢
⎢
⎢
⎢
⎢
⎣
49
Homer
Marge
Bart
Lisa
Maggie
Homer
Marge
Bart
Lisa
Maggie
Homer
Marge
Bart
Lisa
Maggie
Homer
Marge
Bart
Lisa
Maggie
Marge
Marge
Bart
Lisa
Maggie
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
→
1
1
0
1
1
0
0
0
1
0
0
0
1
0
0
1
0
0
0
0
0
1
0
0
0
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
Figure 3.1: The denotation of loves
50
Functional Application
the function and the constituent on the left was the argument. We
also used it to put together a transitive verb with its object, and
in that case the constituent on the left was the function and the
constitutent on the right was was the argument. So the rule of
Functional Application doesn’t care about order, or the syntactic
category of the elements that it combines; it only cares about semantic type. This strategy is what is called type-driven interpretation (Klein & Sag 1985; see also Bittner 1994).
According to Frege’s conjecture, there is only one principle
that we need to derive the meanings of complex expressions from
the meanings of their parts: Functional Application. Indeed, there
is a great deal that can be done with Functional Application alone.
But in later chapters we will introduce a handful of additional principles for semantic composition.
Exercise 5.
(a) What type is ⟦loves⟧?
(b) What is the value of ⟦loves⟧(Lisa)?
(c) What is the value of ⟦loves⟧(Lisa)(Maggie)?
(d) What is the value of ⟦loves⟧(Maggie)(Lisa)?
(e) Is ⟦loves⟧(Maggie)(Lisa) the denotation of Maggie loves Lisa or
Lisa loves Maggie? Think about which argument the verb will
combine with first.
3.3 Connectives
Just as we can use Functional Application to build up the meaning of a sentence, we can also use it to join sentences together
and form complex sentences. Let us define and as a function that
Functional Application
51
takes two truth values (in succession) and returns 1 if both of the
input truth values are 1, and returns 0 otherwise. We can write this
out as follows:
(7)
⎡
⎢ 1 →[
⎢
⎢
⎢
⟦and⟧ = ⎢⎢
⎢
⎢
⎢ 0 →[
⎢
⎣
⎤
1 → 1
] ⎥⎥
⎥
0 → 0
⎥
⎥
⎥
⎥
1 → 0
⎥
] ⎥⎥
0 → 0
⎦
The function that and denotes takes as an argument a truth
value (something of type t ). What it returns is another function
that takes a truth value (type t ) as an argument, and returns a
truth value (type t ). The function that it returns is of type ⟨t , t ⟩.
So the denotation of and is of type ⟨t , ⟨t , t ⟩⟩.
For or, we can have a similar function, one which takes two
truth values and returns 1 if at least one of them is 1.
(8)
⎡
⎢ 1 →[
⎢
⎢
⎢
⟦or ⟧ = ⎢⎢
⎢
⎢
⎢ 0 →[
⎢
⎣
⎤
1 → 1
] ⎥⎥
⎥
0 → 1
⎥
⎥
⎥
⎥
1 → 1
⎥
] ⎥⎥
0 → 0
⎦
This corresponds to the “inclusive” reading of disjunction, as opposed to the “exclusive” reading, which excludes the possibility
that both conjuncts are true. More on this in the next chapter.
In order to include sentences with and and or among the wellformed expressions of our language, we will need to introduce a
new syntax rule. Let us treat and as a word of syntactic category
J (for “junct”, as in “conjunction” and “disjunction”), and phrases
headed by and as being of category JP, and add a rule that allows
a sentence to consist of a sentence followed by a JP.
While we are at it, let us add a rule that lets us negate sentences, and a lexical entry. Since the syntax of negation in English
is a bit complicated, let us just assume that It is not the case that is
a lexical item of category Neg that can be put before a sentence.
52
(9)
Functional Application
Syntax
S
→
VP →
S
→
JP →
S
→
NP VP
V (NP)
S JP
JS
Neg S
This grammar generates analysis trees like the following:
(10)
S
S
JP
NP
VP
Bart
V
J
S
and
snores
NP
VP
Maggie
V
snores
(11)
S
Neg
It is not the case that
S
NP
Bart
VP
V
NP
loves
Maggie
Exercise 6. Annotate the tree in (10) with the semantic types at
each node.
Functional Application
53
3.4 Tiny Fragment of English
Let us summarize the tiny fragment of English that we have specified so far. It has the following basic expressions:
(12)
Lexicon
NP: Bart, Maggie, Homer, Marge, Lisa
V: snores, loves
J: and, or
Neg: it is not the case that
The following context free grammar generates the well-formed syntactic structures of our toy language. Again, the parentheses around
the NP in the VP rule indicate optionality. Notice that there are
several ways of producing an S: with an NP and a VP, with another
S followed by a JP, and with negation preceding another S.
(13)
Syntax
S
→
VP →
S
→
JP →
S
→
NP VP
V (NP)
S JP
JS
Neg S
The expressions of category S are the sentences of the language.
Exercise 7. Which of the following expressions are sentences in
our fragment?
(a) Bart snores loves.
(b) Bart Simpson snores.
(c) Bart Simpson loves.
(d) Bart loves.
(e) Bart snores.
54
Functional Application
(f) It is not the case that Bart snores and Homer loves.
(g) Bart and Homer snore.
Now for the semantics. Remember not to confuse individuals like Maggie, Bart and Homer with names like Maggie, Bart,
and Homer. The names are expressions of the object language
for which we are giving a semantics, and the individuals are living, breathing people. Well, they are cartoon characters but you
know what I mean.
1. ⟦Bart ⟧ = Bart
2. ⟦Maggie⟧ = Maggie
3. ⟦Lisa⟧ = Lisa
4. ⟦Marge⟧ = Marge
5. ⟦Homer ⟧ = Homer
⎡ Bart
→ 1 ⎤⎥
⎢
⎢
⎥
6. ⟦snores⟧ = ⎢ Maggie → 1 ⎥
⎢
⎥
⎢ Homer → 0 ⎥
⎣
⎦
⎡ Bart
⎡
⎢
⎢
⎢
⎢ Bart
→ ⎢ Maggie
⎢
⎢
⎢
⎢ Homer
⎢
⎣
⎢
⎢
⎡ Bart
⎢
⎢
⎢
⎢
7. ⟦loves⟧ = ⎢ Maggie → ⎢ Maggie
⎢
⎢
⎢ Homer
⎢
⎣
⎢
⎢
⎡ Bart
⎢
⎢
⎢
⎢ Homer → ⎢⎢ Maggie
⎢
⎢
⎢
⎢ Homer
⎣
⎣
→
→
→
→
→
→
→
→
→
0
0
1
0
0
0
1
0
1
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
Functional Application
55
⎡
⎤
⎢ 1 →[ 1 → 1 ] ⎥
⎢
⎥
⎢
⎥
0 → 0
⎢
⎥
⎥
8. ⟦and⟧ = ⎢
⎢
⎥
⎢
⎥
1 → 0
⎢
⎥
⎢ 0 →[
⎥
]
⎢
⎥
0
→
0
⎣
⎦
⎡
⎤
⎢ 1 →[ 1 → 1 ] ⎥
⎢
⎥
⎢
⎥
0 → 1
⎢
⎥
⎢
⎥
9. ⟦or ⟧ = ⎢
⎥
⎢
⎥
1 → 1
⎢
⎥
⎢ 0 →[
⎥
]
⎢
⎥
0
→
0
⎣
⎦
10. ⟦not ⟧ = [
1 → 0
]
0 → 1
Composition rules.
1. Non-branching Nodes (NN)
If α is a tree whose only subtree is β, then:
⟦α⟧=⟦β⟧.
2. Functional Application (FA)
If γ is a tree whose only two subtrees are α and β, and ⟦β⟧ is
a function whose domain contains ⟦γ⟧, then:
⟦α⟧= ⟦β⟧(⟦γ⟧).
(Note that the order of β and γ in the tree doesn’t matter.)
Exercise 8. Give a detailed analysis of Lisa loves Maggie based on
this fragment. Begin by drawing the syntax tree. Then compute
the value of each subtree until you have computed the semantic
value of the sentence (which should be a truth value, either 1 or
0).
56
Functional Application
Models, entailment, and truth conditions. At this point, the reader
might have the following worry: The meaning of the word snores
is not (the characteristic function of) any particular set; the meaning of the word snores is something more general that should allow
us to pick out the set of snorers in any given situation. A related
worry is that we haven’t said anything that allows us to explain
entailments, depite the fact that we said in the first chapter that
accounting for entailments is perhaps the primary job of semantics. Both of these worries can be at least partially addressed by
bringing the notion of a model into the picture.
Recall that φ entails ψ if and only if: whenever φ is true, ψ is
true. In setting up the foundations for a semantic theory, we have
to decide exactly what we mean by “whenever”. One possible way
of doing that is with models: In all models where φ is true, ψ is
true. This is typically how entailment is defined in logic. Denotation is relativized to a model, so that instead of ⟦φ⟧ we have ⟦φ⟧M .
The model determines the denotations of words like snores and
loves in the model (sometimes called “content words”, which correspond to the “non-logical constants” in logic).
A sentence might be true in one model but false in another.
For example, take two models M 1 and M 2 , and imagine that Bart
snores in M 1 but doesn’t snore in M 2 . In particular, suppose that
the denotation of snores in M 1 and M 2 is as follows:
⎡
⎢
⎢
• ⟦snores⟧ = ⎢
⎢
⎢
⎣
⎡
⎢
⎢
• ⟦snores⟧M2 = ⎢
⎢
⎢
⎣
M1
Bart
Maggie
Homer
Bart
Maggie
Homer
→ 1 ⎤⎥
⎥
→ 1 ⎥
⎥
→ 0 ⎥⎦
→ 0 ⎤⎥
⎥
→ 0 ⎥
⎥
→ 0 ⎥⎦
Then it will turn out that the denotation of Bart snores in M 1 will
be 1 (true), and its denotation in M 2 will be 0 (false).
If a sentence φ is true in model M , that is, ⟦ψ⟧M = 1, then we
can write M ⊧ φ. Then we can say that φ semantically entails ψ
Functional Application
57
if an only if, for all M , if M ⊧ φ then M ⊧ ψ. Logical entailment
between two sentences is written with the same symbol, so φ ⊧ ψ
stands for “φ semantically entails ψ”.
(14)
Semantic entailment
φ ⊧ ψ iff for all M , if M ⊧ φ then M ⊧ ψ.
For example, given a suitable analysis of someone, it should turn
out that Bart snores semantically entails someone snores, because
in every model where Bart snores is true, someone snores is also
true.
Take another example. We said above that Mary is a cellist entails Mary is a musician. Given our definition of entailment, this
means that in every model where Mary is a cellist is true, Mary is
a musician is also true. So there should be no models where Mary
is a cellist but not a musician. In other words, it should be impossible for Mary to be a cellist without also being a musician. One
could debate whether the entailment really holds in this sense,
especially in light of certain contemporary “music”. But assuming it does hold, then we need to say something about what it
means to be a cellist in order to rule out models where there are
non-musician cellists. Richard Montague suggested doing this via
meaning postulates, which act as restrictions on possible models.
By specifying what a model would have to be like in order for a
sentence to be true, we are effectively assigning truth conditions
to sentences. Recall that knowing the meaning of a sentence does
not require knowing whether the sentence is in fact true; it only
requires being able to discriminate between situations in which
the sentence is true and situations in which the sentence is false.
The truth conditions are the conditions under which the sentence
is true. Another way of thinking about truth conditions is in terms
of models: the truth conditions determine what a model would
have to be like in order for the sentence to be true.
How can we assign truth conditions to sentences of natural
languages like English? Montague’s idea was to treat English is a
formal language, like the languages of formal logic. He said:
58
Functional Application
I reject the contention that an important theoretical
difference exists between formal and natural languages.
... In the present paper I shall accordingly present a
precise treatment, culminating in a theory of truth, of
a formal language that I believe may reasonably be
regarded as a fragment of ordinary English. ... The
treatment given here will be found to resemble the
usual syntax and model theory (or semantics) [due
to Tarski] of the predicate calculus, but leans rather
heavily on the intuitive aspects of certain recent developments in intensional logic [due to Montague himself ]. (Montague 1970b, p.188 in Montague 1974)
What we have done in this chapter so far is very much inspired by
the logical tradition. But to really understand what Montague is
talking about, we must look at logic per se.
4
∣
Logic
4.1 Propositional calculus
We will begin by looking at a simple logic, called propositional
logic or propositional calculus, which contains letters standing for
propositions. Propositions are the kinds of things that can be true
or false, such as It is raining and Snow is white, typically encoded
in English using complete declarative sentences. In propositional
logic, we have expressions like this:
P ∧Q
P ∨Q
P →Q
¬P
‘P and Q’
‘P or Q’
‘P implies Q’
‘not P ’
Propositional calculus is a language, and as such it has a syntax
and a semantics. The expressions just listed are formulas in this
language, which have denotations in a model. In propositional
calculus, the denotations are very simple, always 0 (false) or 1
(true). A model for propositional calculus determines the value
of the proposition letters (e.g. P and Q), and the truth or falsity of
these sentences depends on the value of the proposition letters in
the model and the way that the connectives (∧, ∨, →, and ¬) are
defined. The semantics of the connectives is often laid out using
truth tables, to which we turn first.
59
60
Logic
4.1.1 Truth tables
Imagine you are playing a card game where you’re dealt three cards
from a triple deck (three decks of standard playing cards). Here is
the rule of the game: You win a point if all three of the cards are
of the same suit (the suits being Hearts, Clubs, Diamonds, and
Spades) or all of the same rank (e.g. 3, 4, Ace, King, etc.). Here are
two winning hands:
3
♠
q
3
♠
♠
q
q
q
♠
♠
3
3
3
r
q
r
r
3
r
r
r
Kr
r
r
r
r
4
3
4r
rr
r
r
K
r
rr
3
r
Since it is a triple deck, it could happen that all three cards have
the same rank and the same shape. For example, you could get
three Aces of Spades:
A
A
♣
A
♣
♣
♠
♠
q
q
A
♠
q
A
A
Under a normal interpretation of the rule of the game as it was
stated (“You win a point if all of the cards are of the same suit or
all of the same rank”), you would also win in this situation. Let’s
express the rule using symbols to bring out its structure.
Logic
61
R = You get three of the same rank.
S = You get three of the same suit.
The rule of the game was that you win if R is true or S is true. An or
statement is called a disjunction. Under the interpretation where
you do win when both are true, the disjunction in this case includes the possibility that both of the disjuncts are true.
Let us represent “you get three of the same rank or three of the
same suit” as:
R ∨S
The ∨ symbol represents “or”, i.e. disjunction. We can represent
the meaning of ∨ using a truth table:
R
S
R ∨S
True
True
False
False
True
False
True
False
True
True
True
False
This table represents inclusive disjunction, because it includes the
case where both of the so-called disjuncts are true.
Exclusive disjunction, where only one of the disjuncts can
(and must) be true, is defined like this:
R
S
R ⊕S
True
True
False
False
True
False
True
False
False
True
True
False
This is closer to the intuitive meaning of or in I’ll see you today
or tomorrow. One might imagine that natural language or is ambiguous between inclusive and exclusive disjunction, but there is
reason to believe that inclusive reading is what or really denotes,
62
Logic
and that the exclusive reading arises via a conversational implicature in certain contexts. One argument for this comes from negation: If I say I won’t see you today or tomorrow, it means that I will
not see you today and I won’t see you tomorrow. We can get this
interpretation by negating the inclusive interpretation (‘it is not
the case that at least one of the disjuncts is true’). Conversational
implicatures of the kind that would be involved here (‘scalar implicatures’) typically disappear under negation, so this fact is easily explained under the implicature analysis. Under the ambiguity
analysis, it is not clear why an exclusive reading should disappear
under negation. (There is of course much more to say on this issue.)
Now suppose the game gets harder: you have to get all the
same suit and all the same rank in order to win. Then you would
need for R and S to be true. In other words, the conjunction of R
and S has to hold. Conjunction can be represented as follows:
R ∧S
The truth table for conjunction is as follows:
R
S
R ∧S
True
True
False
False
True
False
True
False
True
False
False
False
So far, we have discussed three different binary connectives:
inclusive disjunction, exclusive disjunction, and conjunction. The
truth of formulas containing these connectives depends on the
truth values of the formulas it joins together. These are thus truthfunctional connectives.
Another binary truth-functional connective is material implication, which corresponds roughly to conditional sentences of
the form if A then B. The truth table for this one is a bit tricker
Logic
63
to understand intuitively, but here goes. Let W = you win. We said
above that you win if and only if you get all the same rank or all the
same suit. And, in particular, the following should hold, according
to the rules of the game:
If you get all the same rank, then you win.
We can symbolize this as follows:
R →W
Suppose that an inspector comes along and wants to make sure
that the game is operating according to the rules. In particular, he
wants to make sure that R → W holds. There are several cases to
check:
• You get all the same rank and you win. ✓
• You get all the same rank and you don’t win. Uh oh!
• You don’t get all the same rank but you still win (presumably
you got all the same suit). ✓
• You don’t get all the same rank and you don’t win (presumably you didn’t get all the same suit). ✓
The only situation that falsifies R → W is the one where R is true,
and W is false. This is reflected in the truth table for R → W :
R
W
R →W
True
True
False
False
True
False
True
False
True
False
True
True
The expression on the left-hand side of the arrow (R, here) is the
antecedent and the expression on the right-hand side (W , here)
is the consequent. So in general, a conditional sentence is false if
64
Logic
the antecedent is true and the consequent is false, and true otherwise.
While it seems intuitively clear that a conditional is false when
the antecedent is true and the consequent is false, it admittedly
seems less intuitively clear that a conditional is true in the circumstance where the antecedent is false. For example, the moon
is not made of green cheese. Does that mean that If the moon is
made of green cheese, then I had yogurt for breakfast this morning
is true? Intuitively not. In English, at least, conditionals are used
to express regularities, so one might reasonably argue that they
cannot be judged as true or false in a single situation. In order
to capture the meaning of English conditionals, we need somewhat more sophisticated technology. (The interested reader is referred to the work of David Lewis, Robert Stalnaker, and Angelika
Kratzer, among many others.) But if we are forced to identify the
truth-functional connective that most closely resembles if then,
we will have to choose material implication, as material implication is the truth-functional connective that corresponds most
closely to English if-then.
The truth table for negation, written ¬, is simple, because negation is a unary connective. If R is true, then ¬R is false. If R is false,
then ¬R is true:
R
¬R
True
False
False
True
Let us consider when the negation of R ∧ S is true. To find out, we
first find out when R ∧ S is true, and then apply negation to that.
R
S
R ∧S
¬[R ∧ S ]
True
True
False
False
True
False
True
False
True
False
False
False
False
True
True
True
Logic
65
Note: We are using the brackets [ ] to show that we are applying
negation to the conjunction of R and S, rather than R.
4.1.2 Reasoning with truth tables
In the previous chapter, we said that φ entails ψ if and only if:
Whenever φ is true, ψ is true too. That definition works in the
context of propositional logic as well. Each situation corresponds
to a row of the truth table, so we can check whether whether φ
entails ψ by looking at all of the rows of the truth table where φ
holds, and checking whether ψ holds in that row as well.
Exercise 1. Does P entail P ∨ Q? Explain, using a truth table.
The notion of entailment plays a role in determining whether
or not a logical argument is valid. An argument consists of a set
of premises and a conclusion, and it is valid if the conclusion is
true whenever all of the premises are true. (Note that validity is a
property of arguments, while truth is a property of propositions,
and not vice versa.) In other words, it is valid if the premises entail
the conclusion.
Here is an example of a valid argument.
(1)
R →W
R
W
(Premise 1)
(Premise 2)
(Conclusion)
This is an instance of the form of reasoning called modus ponens.
We can show that it is valid using a truth table. We need a column for both of the premises and the conclusion, and then we
can check that the argument is valid by determining whether the
conclusion is true whenever the all of the premises are true.
66
Logic
R (Premise)
W (Conclusion)
R → W (Premise)
True
True
False
False
True
False
True
False
True
False
True
True
As this table shows, there is only one situation in which R and
[R → W ] are both true. In that situation, W is true. Hence the
premises entail the conclusion and the argument is valid.
Exercise 2. Use a truth table to show that the following argument is not valid. Point out the counterexample, where all of the
premises are true, and yet the conclusion is false.
(2)
R →W
¬R
¬W
(Premise 1)
(Premise 2)
(Conclusion)
We can also use truth tables to show that two expressions are
equivalent, i.e., true under exactly the same circumstances. For
example, P → Q is equivalent to [¬P ∨ Q ]. We can show this using
a truth table. By making columns for both expressions and comparing them, we can see that they are true under exactly the same
circumstances.
P
Q
P →Q
¬P
[¬P ∨ Q ]
True
True
False
False
True
False
True
False
True
False
True
True
False
False
True
True
True
False
True
True
Logic
67
Exercise 3. Show that P → Q is also equivalent to ¬Q → ¬P by filling in the truth table below.
P
True
True
False
False
Q
True
False
True
False
¬P
¬Q
¬Q → ¬P
False
False
True
True
A tautology is an expression that is true in every situation. So
you can tell whether an expression is a tautology by looking at the
pattern of Trues and Falses in the column underneath it in a truth
table: If they’re all true, then it is a tautology. Here is a tautology:
P ∨ ¬P (e.g. It is raining or it is not raining):
P
¬P
P ∨ ¬P
True
False
False
True
True
True
Exercise 4. Which of the following are tautologies?
(a) P ∨ Q
(b) [P → Q ] ∨ [Q → P ]
(c) [P → Q ] ↔ [¬Q ∨ ¬P ]
Support your answer with truth tables.
Two logical expressions are contradictory if for every assignment of values to their variables, their truth values are different.
For example P and ¬P are contradictory.
68
Logic
P
¬P
True
False
False
True
Another contradictory pair is P → Q and P ∧ ¬Q.
P
Q
P →Q
¬Q
P ∧ ¬Q
True
True
False
False
True
False
True
False
True
False
True
True
False
True
False
True
False
True
False
False
4.1.3 Grammar of PL
Propositional Logic is a language (or a class of languages), and
languages have both grammar and semantics. The grammar specifies the well-formed formulas of the language. The semantics
specifies the semantic value of every well-formed formula, given
a model. In other words, the semantics assigns a denotation to
every well-formed formula in a model.
It is sometimes not understood that logics are languages. Some
authors suggest that logic is just a way of representing semantics.
But logic is not just semantics – logics have grammar and semantics.
Let us now define the grammar for a propositional language
including the connectives described in the previous section. We
will call the language PL. We use the following rules for defining
what counts as a formula of PL.
(3)
Syntax of PL
1. Atomic formulae
Each of the proposition letters P , Q, R, and S is a formula (an atomic formula in particular).
Logic
69
2. Negation
If φ is a formula, then ¬φ is a formula.
3. Binary connectives
If φ and ψ are formulas, then so are:
(a) [φ ∧ ψ]
(b) [φ ∨ ψ]
(c) [φ → ψ]
(d) [φ ↔ ψ]
In these rules, the Greek letters φ ‘phi’ and ψ ‘psi’ are variables
ranging over formulas of the language PL. These symbols re placeholders for expressions of the language, not actual expressions of
the language. In this case, PL is the object language – the language
we are studying – and English is the meta-language – the language
we are using to talk about the object language. The Greek letters are variables in the meta-language, so we can refer to them
as meta-language variables.
The syntax rules for PL allow us to construct an infinite number of formulas. For example, we can start with P , and then apply
the second rule (for negation) to produce the formula ¬P . Since
that is a formula, we can apply the same rule again and produce
¬¬P . And then ¬¬¬P . And so on and so forth.
Note that brackets are typically omitted at the outermost edges
of a formula, but technically, they are present in every formula
containing a binary connective. (The rule for negation does not
introduce brackets.) In some cases, these brackets are crucial for
understanding the structure. For example, [¬P ∧ Q ] means something very different from ¬[P ∧ Q ].
Of course, not every string using these symbols is a formula.
For example:
[¬][∧pq →
is not a formula, because it is not licensed by the syntax rules.
70
Logic
4.1.4 Semantics of PL
In giving semantic sto the expressions of propositional logic, we
implement Frege’s (1948) idea that the denotation of a sentence is
a truth value. In his famous article Über Sinn und Bedeutung (‘On
Sense and Reference’), he writes (p. 216): “By the truth value of
a sentence I understand the circumstance that it is true or false.
There are no further truth values. For brevity I call the one the
true, the other the false.” We use the number 1 to represent “the
true” and the number 0 to represent “the false”. These are the two
possible semantic values for a formula:
• 1 (True)
• 0 (False)
(Sometimes people use three truth values, including an undefined
value, or four, adding an over-defined value, or more.)
The truth of a formula depends on the state of the world. This
is encoded using a model (sometimes also called a structure),
which determines what is and is not the case. (The notion of
model in this sense is due to the logician Alfred Tarski. Richard
Montague brought the use of models into semantics for natural
language, popularizing what is known as model-theoretic semantics.) We will see more complicated models for other logics later
on, but a model for propositional logic is just a function that takes
a proposition letter and returns 1 or 0. For example, it might be
the case in model M 1 that P is true, Q is false, R is true and S is
false.
⎡ P → 1 ⎤
⎢
⎥
⎢ Q → 0 ⎥
⎢
⎥
⎥
M1 = ⎢
⎢ R → 1 ⎥
⎥
⎢
⎢ S → 0 ⎥
⎣
⎦
Thus M 1 (P ) = 1, M 1 (Q ) = 0, M 1 (R ) = 1 and M 1 (S ) = 0. Later on,
models will become more complicated, but here, a model is just a
function that takes a proposition letter and returns a 1 or 0.
Logic
71
The semantics of a truth-functional connective (→, ∧, ∨) is
defined in terms of the truth values of the formulas it joins together. Each possible assignment of truth values to propositional
variables (each different model) is represented by a different row
of the truth table. The following rules are another way of expressing what we said using truth tables above.
(4)
Semantics of PL
1. Atomic formulae
M
If φ is an atomic formula, then ⟦φ⟧ = M (φ)
2. Negation
⟦¬φ⟧M = 1 if and only if ⟦φ⟧M = 0
3. Binary connectives
= 1 if and only if ⟦φ⟧M = 1 and ⟦ψ⟧M = 1
M
M
(b) ⟦φ ∨ ψ⟧ = 1 if and only if ⟦φ⟧ = 1 or ⟦ψ⟧ = 1
M
M
M
(c) ⟦φ → ψ⟧ = 1 if and only if ⟦φ⟧ = 0 or ⟦ψ⟧ = 1
M
M
M
(d) ⟦φ ↔ ψ⟧ = 1 if and only if ⟦φ⟧ = ⟦ψ⟧
(a) ⟦φ ∧ ψ⟧
M
M
Exercise 5. Specify a model M 2 for PL in which [P → Q ] is false,
and another model M 3 in which it is true, and explain why the
formula is false in M 2 but true in M 3 .
4.2 First-order logic without variables (L0 )
We ultimately want to be able to derive the fact that, for example,
Beth speaks French implies Beth speaks a European language. In
order to do this, we will need more than just connectives; we need
to break the formulas down into smaller parts. In this section, we
will describe a logic that has something akin to verbs like speak
(predicates) and proper names like Beth (terms).
72
Logic
4.2.1 Syntax of L0
In this section, we will define the grammar of a logical language,
which we will call L0 . This consists in a set of rules specifying, for
each syntactic category, what counts as a well-formed expression
of that category. The syntactic categories include:
• terms (names for individuals)
• predicates (names for properties and relations)
• formulas (sentences)
4.2.1.1 Syntax
The syntax of L0 is as follows. Again, we are using Greek letters (π
‘pi’, α ‘alpha’, φ ‘phi’, and ψ ‘psi’) for variables over expressions of
the object language (meta-language variables).
(5)
Syntax of L0
1. Basic expressions
(a) individual constants: D, N , J , M
(b) unary predicates: HAPPY, BORED, LINGUIST, PHILOSO PHER , CELLIST, SMOKES
(c) binary predicates:
KISSED, LOVES , ADMIRES
If α is an individual constant, then α is a term.
2. Predication
(a) If π is a unary predicate and α is a term, then π(α)
is a formula.
(b) If π is a binary predicate and α is a term, then
π(α) is a unary predicate.1
1
This bit is non-standard for first-order logic. This clause is normally given
as “If π is a binary predicate and α and β are terms, then π(α, β) is a formula.”
We are doing things differently because we are treating transitive verbs like the
binary predicates of logic, and transitive verbs combine first with their direct
Logic
73
3. Negation
If φ is a formula, then ¬φ is a formula.
4. Binary connectives
If φ is a formula and ψ is a formula, then so are [φ∧ψ],
[φ ∨ ψ], [φ → ψ], and [φ ↔ ψ].
Let us use the first rule to create a complex expression. We
need a unary predicate and a term. Let’s use HAPPY and M. Since
HAPPY is a unary predicate and M is a term, HAPPY ( M ) is a formula.
With the rule for binary connectives, therefore, we can derive that
the following is a formula:
[HAPPY(M) ↔ HAPPY(M)]
Notice the square brackets: The syntactic composition rules specifically mention them, so they are part of the formula.
Now look at the second predication rule (“If π is a binary predicate and α is a term, then π(α) is a unary predicate”). This rule
says that if you have a binary predicate, and you combine it with
a term, then you end up with a unary predicate. Take the binary
predicate LOVES. Combined with the term M, we get LOVES(M),
which is a unary predicate, according to this rule. We can take
such an expression and combine it with a term to form a formula.
Since LOVES(M) is a unary predicate and J is a term, LOVES(M)(J) is
a formula.
Exercise 6. Which of the following are formulas of L0 ?
(a) ¬¬HAPPY(M)
(b)
HAPPY ( A )
object and then with the subject. Just as a transitive verb plus its object is of the
same syntactic category as an intransitive verb, a binary predicated plus its first
argument is of the same category as a unary predicate here.
74
Logic
(c)
HAPPY ( M )( M )
(d)
LOVES ( J )( M ) ∨ LOVES ( M , J )
(e) [LOVES(J)(M) ↔ LOVES(M)(J)]
(f) [KISSED(M) ↔ LOVES(M)(J)]
(g) ¬[LOVES(J)(M) ← ¬LOVES(M)(J)]
4.2.2 Semantics of L0
4.2.2.1 Models
The semantics of L0 will be given in terms of a model. The model
represents a possible state of affairs. It consists of a domain of
individuals D and an interpretation function I which specifies
the values of all of the basic expressions in the lexicon.
Let us define a model M 1 = ⟨D 1 , I 1 ⟩.2 The domain of this model
consists of three individuals: Maggie, Bart, and Homer Simpson.
(6)
Domain of M 1
D 1 = {Maggie Simpson, Bart Simpson, Homer Simpson}
The interpretation function I 1 will assign values to all of the expressions in the lexicon, including terms, unary predicates, and
binary predicates. The interpretation of each of the terms will be
an individual in D 1 .
(7)
2
Interpretation of terms in M 1
a.
I 1 (D) = Homer
b.
I 1 (N) = Bart
c.
I 1 (J) = Bart
The notation ⟨a, b ⟩ represents an ordered pair whose first member is a and
whose second member is b.
Logic
75
d.
I 1 (M) = Maggie
The interpretation of a unary predicate will be a function that
takes an individual as input and spits back either 1 (“true”) or 0
(“false”). For example, we could assume that the interpretation of
the constant BORED maps Maggie and Homer to “false”, and Bart
to “true”.
(8)
⎡ Bart
→ 1 ⎤⎥
⎢
⎥
⎢
I 1 (BORED) = ⎢ Homer → 0 ⎥
⎥
⎢
⎢ Maggie → 0 ⎥
⎦
⎣
For the unary predicate
following function:
(9)
HAPPY ,
let us assume that it denotes the
⎡ Bart
→ 1 ⎤⎥
⎢
⎢
⎥
I 1 (HAPPY) = ⎢ Homer → 0 ⎥
⎢
⎥
⎢ Maggie → 1 ⎥
⎣
⎦
Exercise 7.
(a) What does I 1 (HAPPY) give as output when given Bart as an
input?
(b) In other words, what is I 1 (HAPPY)(Bart)?
(c) In other words, is Bart happy in M 1 ?
Exercise 8. True or false:
(a) I 1 (BORED)(Bart) = 1
(b) I 1 (BORED)(Bart) = I 1 (HAPPY)(Bart)
(c) I 1 (N) = Bart
76
Logic
(d) I 1 (BORED)( I 1 (N)) = 1
(e) I 1 (N) = I 1 (J)
Now for the fun part. Binary predicates will denote functions
which, when given an individual, return a another function. These
functions are by definition of type ⟨e, ⟨e, t ⟩⟩ (input e, output ⟨e, t ⟩).
For example, given the individual Maggie, the following function
returns another function which, given the individual Maggie again,
returns true.
(10)
⎡ Maggie
⎡
⎢
⎢
⎢ Maggie → ⎢⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎢
⎣
⎢
⎢
⎡ Maggie
⎢
⎢
⎢
⎢
I 1 (LOVES) = ⎢ Homer → ⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎣
⎢
⎢
⎡
Maggie
⎢
⎢
⎢
⎢
⎢ Bart
→ ⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎣
⎣
→
→
→
→
→
→
→
→
→
1
0
0
0
1
0
0
0
1
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
Let us call this function as f . As you can see, f (Maggie)(Maggie)=
1, but f (Maggie)(Bart)= 0. In general, this corresponds to the ‘equal’
relation: It returns true only when both arguments are the same.
This describes a situation in which each individual loves themself,
and nobody else.
Exercise 9.
(a) What is the value of I 1 (LOVES) applied to Maggie? Hint: It is
a function. Specify the function as a set of input-output pairs
(using either set notation or arrow notation).
Logic
77
(b) What is the value of that function (the one you gave in your
previous answer) when applied to Maggie?
(c) In other words, what is I 1 (LOVES)(Maggie)(Maggie)?
(d) So, does Maggie love herself in M 1 ?
Let us suppose that LOVES denotes this ‘self’ function and KISSED
denotes the ‘other’ function, which returns 1 if and only if the two
arguments are different.
⎡ Maggie
⎡
⎢
⎢
⎢ Maggie → ⎢⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎢
⎣
⎢
⎡ Maggie
⎢
⎢
⎢
⎢
⎢
I 1 (KISSED) = ⎢ Homer → ⎢ Homer
⎢
⎢
⎢ Bart
⎢
⎣
⎢
⎢
⎡ Maggie
⎢
⎢
⎢
⎢
⎢ Bart
⎢ Homer
→
⎢
⎢
⎢
⎢ Bart
⎣
⎣
→
→
→
→
→
→
→
→
→
0
1
1
1
0
1
1
1
0
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
Let us assume that the first individual that the function combines
with is the person being kissed, and the second individual is the
kisser, because a transitive verb combines first with its object nad
then with its subject. Then this denotation for kissed means that
Bart and Homer kissed Maggie, Maggie and Bart kissed Homer,
and Maggie and Homer kissed Bart. Applied to Maggie, this function yields the function that returns 1 for Bart and Homer:
⎡ Maggie → 0 ⎤
⎢
⎥
⎢
⎥
I 1 (KISSED)(Maggie) = ⎢ Homer → 1 ⎥
⎢
⎥
⎢ Bart
→ 1 ⎥⎦
⎣
And I 1 (KISSED)(Maggie)(Homer) = 1. Now we have a complete
specification of all of the values for the basic expressions in the
logic.
78
Logic
Exercise 10. True or false: For all individuals k in D e (namely Maggie, Homer, and Bart), I 1 (LOVES)(k )(k ) = 1. In other words, everybody loves him- or herself in M 1 .
Exercise 11.
4.2.2.2 Composition rules
So far we have talked about the basic expressions of the language.
What about the complex expressions of the language? How do we
M
get semantic values for them? We use a valuation function ⟦ ⋅ ⟧ ,
which is built up recursively3 on the basis of the basic interpretation function I . The valuation function assigns to every expression α of the language (not just the basic expressions) a semantic
M
value ⟦α⟧ .
(11)
Semantics of L0
1. Basic expressions
If α is a basic expression, then:
⟦α⟧M = I (α).
where I is the interpretation function of the model M .
(Recall that a model M consists of a domain D and an
interpretation function I .)
2. Predication
3
A recursive definition is one that uses itself as part of the definition (but is
not circular). For example, we could define the set of natural numbers recursively by saying: “0 is a number (this is the base case); and if n is a number, then
n + 1 is a number.” The definition of ⟦ ⋅ ⟧ is recursive because the denotation of a
complex expression is defined in terms of the denotations of its parts.
Logic
79
If π is a predicate (unary or binary)4 and α is a term,
then:
⟦π(α)⟧M = ⟦π⟧M (⟦α⟧M ).
3. Negation
⟦¬φ⟧M = 1 if and only if ⟦φ⟧M = 0.
4. Binary Connectives
(a)
(b)
(c)
(d)
⟦φ ∧ ψ⟧M = 1 if and only if ⟦φ⟧M = 1 and ⟦ψ⟧M = 1.
⟦φ ∨ ψ⟧M = 1 if and only if ⟦φ⟧M = 1 or ⟦ψ⟧M = 1.
⟦φ → ψ⟧M if and only if ⟦φ⟧M = 0 or ⟦ψ⟧M = 1.
⟦φ ↔ ψ⟧M if and only if ⟦φ⟧M = ⟦ψ⟧M .
Notice that the value of a basic expressions depends on the model,
while the contribution of a connective to the meaning is the same
regardless of the model. In this sense, the predicates and individual constants are non-logical constants, while the connectives
are logical constants. (We refer to both as “constants” in order
to distinguish them from variables, which we will introduce later
on.) Punctuation symbols that are used for grouping like brackets
are not typically considered constants.
Exercise 12. For each of the following, say whether it is a logical
constant or a non-logical constant or neither.
(a) ∧
(d) →
(g)
(b)
(e) ¬
(h) Bart
D
(c) (
4
(f)
LOVES
HAPPY
(i) ]
(j)
LOVES ( D )
(k) ←
(l) ∨
This is a non-standard way of treating predication in first-order logic. NorM
M
M
mally, unary predicates denote sets, so ⟦π(α)⟧ = 1 iff ⟦α⟧ ∈ ⟦π⟧ . Binary
M
predicates normally denote relations (sets of ordered pairs), so ⟦π(α, β)⟧ = 1
M
M
M
iff ⟨⟦α⟧ , ⟦β⟧ ⟩ ∈ ⟦π⟧ . We are doing things this way instead, because we are
treating binary predicates of logic like transitive verbs in natural languages, and
transitive verbs combine first with their direct object and then with the subject.
80
Logic
Let us use these rules to compute the semantic value of the expression BORED(D) with respect to the model that we have specified, M 1 . Since we are starting with an expression of the form π(α)
where π is a predicate and α is a term, we can use the Predication
rule in (11). This rule tells us that:
⟦BORED(D)⟧M1 = ⟦BORED⟧M1 (⟦D⟧M1 )
M
M
So now we have to compute ⟦BORED⟧ 1 (⟦D⟧ 1 ). To do this, we
M
M
have to compute ⟦BORED⟧ 1 and ⟦D⟧ 1 . The Basic expressions
rule in (11) tells us the following two facts:
⟦BORED⟧M1 = I 1 (BORED)
⟦ D ⟧M 1 = I 1 ( D )
We know that I 1 (BORED) = {⟨Maggie, 0⟩, ⟨Homer, 0⟩, ⟨Bart, 1⟩} and
I 1 (D)=Homer. Now we just have to apply the former to the latter,
in order to get the value of the expression we are trying to evaluate.
I 1 (BORED) is a function that returns 0 when Homer given as in
input. Therefore, in M 1 , the value of the whole expression we are
trying to evaluate, namely BORED(D), is 0. In other words, this
formula is false in M 1 .
Exercise 13. Compute the semantic value of each of the wellformed formulas in Exercise (6) with respect to M 1 .
Exercise 14. Consider a fantasy model M f = ⟨D, I f ⟩ in which everybody is happy:
⎡ Bart
→ 1 ⎤⎥
⎢
⎢
⎥
I f (HAPPY) = ⎢ Homer → 1 ⎥
⎥
⎢
⎢ Maggie → 1 ⎥
⎦
⎣
Logic
What is ⟦HAPPY(D)⟧
81
Mf
?
4.3 First-order logic with variables (L1 )
4.3.1 Syntax of L1
Predicate calculus (a.k.a. first-order logic) famously has sentences
with quantifiers and variables like these:
∀x HAPPY(x )
∃x ¬HAPPY(x )
‘for all x, x is happy’
‘there exists an x such x is not happy’
We ignored variables and quantifiers in the previous section, and
now we will add them to our language. We will call our new language L1 , as it is a version of first-order logic. L1 is a first-order
logic in the sense that the variables range only over individuals,
and not over more complex objects like predicates.5
Before we define the syntax and semantics of the language formally, let us spend some time getting used to the formalism. Take
the following formula:
∀x [LINGUIST(x ) → ∃ y [PHILOSOPHER( y ) ∧ ADMIRES( y )(x )]]
If were to read this aloud symbol for symbol, we would say, “For
every x, if x is a linguist, then there exists a y such that y is a philosopher and x admires y.” (We assume here and throughout that the
first argument of ADMIRES corresponds to the direct object of the
5
The only way in which L1 differs from standard first-order logic is in the way
that predication is dealt with. As in L1 , we write LOVES ( D )( N ) instead of the more
standard LOVES ( N , D ), and the predicates denote functions rather than sets or
relations. This makes our language more like English, where transitive verbs
combine first with their objects and then with their subjects, and also smooths
the transition to Lλ , which will come in §4.4.
82
Logic
verb, and the second object corresponds to the subject of the verb,
because natural language transitive verbs combine with their arguments in that order.) A more natural way of putting this would
be “Every linguist admires a philosopher.” But notice that “Every
linguist admires a philosopher” is actually ambiguous. It could
mean that for every linguist, there is a philosopher that the linguist admires (possibly a different philosopher for every linguist),
or it could mean that there is a philosopher such that every linguist admires that philosopher. The latter reading could be rendered logically as follows:
∃ y [PHILOSOPHER( y ) ∧ ∀x [LINGUIST(x ) → ADMIRES( y )(x )]]
Predicate logic is thus a tool for teasing apart these kinds of ambiguities in natural language. What we have just seen is an instance
of quantifier scope ambiguity. The first reading is the one where
“every linguist” takes wide scope over “a philosopher”. On the second reading, “every linguist” has narrow scope with respect to “a
philosopher”.
Quantifiers can also take wide or narrow scope with respect
to negation. Consider the sentence “everybody isn’t happy”. This
could mean either one of the following:
∀x ¬[HAPPY(x )]
¬∀x [HAPPY(x )]
The one where the universal quantifier takes wide scope over negation says, “for every x, it is not the case that x is happy.” The one
where the quantifier has narrow scope with respect to negation
says, “it is not the case that for every x, x is happy.” The first one
implies that nobody is happy. The second one implies merely that
there is at least one person who is not happy.
Logic
83
Exercise 15. For each of the following formulas, say (i) how you
would read the formula aloud, using phrases like ‘for all x’ and
‘there exists an x such that’ and (ii) give a natural paraphrase in
English.
(a) ∀x
BORED (x )
(b) ∀x [BORED(x ) ∧ HAPPY(x )]
(c) ∃x [BORED(x ) ∧ HAPPY(x )]
(d) ∃x [BORED(x ) ∨ HAPPY(x )]
(e) ∀x [BORED(x ) → HAPPY(x )]
(f) ∀x ¬BORED(x )
(g) ∃x ¬BORED(x )
(h) ¬∃x
BORED (x )
(i) ∀x ∃ y
LOVES ( y )(x )
Exercise 16. For each of the following sentences, say which of the
formulas above it matches (if any). (In some cases, the sentence
might match two formulas.)
(a) Somebody is bored and happy.
(b) Everybody is bored and happy.
(c) Everybody who is bored is happy.
(d) Nobody is bored.
(e) Somebody is not bored.
84
Logic
(f) Somebody is bored or happy.
(g) Everybody loves somebody.
(h) Somebody loves everybody.
Exercise 17. Which of the following statements in first-order logic
better represents the meaning of Every cellist smokes?
(a) ∀x [CELLIST(x ) → SMOKES(x )]
(b) ∀x [CELLIST(x ) ∧ SMOKES(x )]
Now let us start defining the syntax of this language formally.
We will allow an infinite number of variables v 0 , v 1 , v 2 , ... all of type
e, but use the following shorthands:
• x is v 0
• y is v 1
• z is v 2
We will also add new formation rules for the universal quantifier
∀, and the existential quantifier ∃.
(12)
Syntax of L1
1. Basic expressions
The basic expressions of L1 are of the following categories:
(a) individual constants: D, N , J , M
(b) unary predicates: HAPPY, BORED, LINGUIST, PHILOSO PHER , CELLIST, SMOKES
Logic
85
(c) binary predicates:
KISSED, LOVES , ADMIRES
(d) individual variables, v n for every natural number
n (new!)
If α is an individual constant or a variable, then α is a
term.
2. Predication
(a) If π is a unary predicate and α is a term, then π(α)
is a formula.
(b) If π is a binary predicate and α is a term, then
π(α) is a unary predicate.
3. Negation
If φ is a formula, then ¬φ is a formula.
4. Binary connectives
If φ is a formula and ψ is a formula, then so are [φ∧ψ],
[φ ∨ ψ], [φ → ψ], and [φ ↔ ψ].
5. Quantification
(a) If u is a variable and φ is a formula, then ∀uφ is a
formula. (new!)
(b) If u is a variable and φ is a formula, then ∃uφ is a
formula. (new!)
The last two clauses are new; we had all of the other ones in L0 .
And notice that in the first two clauses, the word term applies not
only to constant terms like D but also variable terms like x.
4.3.2 Semantics
We will use the same kind of models for L 1 as we used for L 0 ,
so a model for L 1 is a pair ⟨D, I ⟩ where D is a set of individuals
and I is an interpretation function giving a value to every nonlogical constant. Informally, ∀x HAPPY(x ) ‘for all x, x is happy’
is true in a model M iff: no matter which individual we assign
to x, HAPPY(x ) is true. This means that for all individuals k, it
86
Logic
M
holds that ⟦HAPPY⟧ (k ) = 1. Likewise, informally, ∃x HAPPY(x )
is true iff: we can find some individual to assign to x such that
HAPPY (x ) is true. In other words, there is some individual k such
M
that ⟦HAPPY⟧ (k ) = 1.
In order to interpret formulas with variables, we will make interpretation relative to a model and an assignment function. So
instead of
⟦φ⟧M
we will now write:
⟦φ⟧M ,g
where g stands for an assignment function.
An assignment function assigns individuals to variables. Here
are some examples of assignment functions:
⎡
⎢
⎢
⎢
g1 = ⎢
⎢
⎢
⎢
⎣
x → Maggie
y → Bart
z → Bart
...
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
⎡
⎢
⎢
⎢
g2 = ⎢
⎢
⎢
⎢
⎣
⎤
x → Bart
⎥
y → Homer ⎥
⎥
⎥
⎥
z → Bart
⎥
⎥
...
⎦
The domain of an assignment function is the set of variables (v n
for all n).
The variables are symbols of the object language, and the interpretation of these symbols will depend on the assignment function. The denotation of the variable x with respect to model M
and assignment function g is simply whatever g maps x to. (The
denotation of the variable depends only on the assignment function, and not on the model.) We can express this more formally as
follows:
⟦x ⟧M ,g = g (x )
M ,g 1
For example, ⟦x ⟧
for any model M .
Exercise 18.
= g 1 (x ) =Maggie, and ⟦x ⟧M ,g 2 = g 2 (x ) =Bart
Logic
87
(a) What is g 1 ( y )?
(b) What is ⟦ y ⟧
M ,g 1
(for any model M )?
(c) What is g 2 ( y )?
(d) What is ⟦ y ⟧
M ,g 2
(for any model M )?
M ,g
The rule ⟦x ⟧
= g (x ) is one of the semantic rules for L 1 , one
which we didn’t have in L 0 because we didn’t have variables yet.
All of the the semantic rules that we had for L0 will carry over to L1 ,
except that the denotation is always relative not only to a model
M but also to an assignment g . Furthermore, we will add rules for
universal and existential quantification (to be explained below).
(13)
Semantics of L1
1. Basic expressions
M ,g
(a) If α is a non-logical constant, then ⟦α⟧
= I(α).
M ,g
(b) If α is a variable, then ⟦α⟧
= g (α). (new!)
2. Predication
If π is a unary or binary predicate and α is a term, then
⟦π(α)⟧M .g = ⟦π⟧M ,g (⟦α⟧M ,g ).
3. Negation
⟦¬φ⟧M ,g = 1 if ⟦φ⟧M ,g = 0; otherwise ⟦¬φ⟧M ,g = 0.
4. Binary Connectives
(a)
(b)
(c)
(d)
⟦φ ∧ ψ⟧M ,g = 1 iff ⟦φ⟧M ,g = 1 and ⟦ψ⟧M ,g = 1.
⟦φ ∨ ψ⟧M ,g = 1 iff ⟦φ⟧M ,g = 1 or ⟦ψ⟧M ,g = 1.
⟦φ → ψ⟧M ,g = 1 iff ⟦φ⟧M ,g = 0 or ⟦ψ⟧M ,g = 1.
⟦φ ↔ ψ⟧M ,g = 1 iff ⟦φ⟧M ,g = ⟦ψ⟧M ,g .
5. Quantification (new!)
(a) ⟦∀vφ⟧
M ,g
= 1 iff for all individuals k ∈ D:
⟦φ⟧M ,g [v ↦k ] = 1
88
Logic
M ,g
(b) ⟦∃vφ⟧
= 1 iff there is an individual k ∈ D such
that that:
⟦φ⟧M ,g [v ↦k ] = 1.
Don’t worry if you don’t understand all of these yet. If you get to
the end of the chapter and you still don’t understand, then you
can start worrying.
There is a piece of notation in the semantic rules for existential and universal quantification that we haven’t explained. Let us
begin by explaining that. The expression
g [u ↦ k ]
is used to describe an assignment function that is exactly like g
save that g (u ) = k. If g already maps u to k then, g [u ↦ k ] is the
same as g . This lets us keep everything the same in g except for
the variable of interest. For example, using g 1 from above,
⎡
⎢
⎢
⎢
g1 = ⎢
⎢
⎢
⎢
⎣
x → Maggie
y → Bart
z → Bart
...
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
g 1 [ y ↦ Homer] would be as follows:
⎡
⎢
⎢
⎢
g 1 [ y ↦ Homer] = ⎢
⎢
⎢
⎢
⎣
x → Maggie ⎤
⎥
y → Homer ⎥
⎥
⎥
⎥
z → Bart
⎥
⎥
...
⎦
We changed it so that y maps to Homer and kept everything else
the same.
Exercise 19.
Logic
89
(a) What is g 1 [z ↦ Homer](x )? (I.e., what does g 1 [z ↦ Homer
assign to x?)
(b) What is g 1 [z ↦ Homer]( y )?
(c) What is g 1 [z ↦ Homer](z )?
To see what this ability to modify assignments does for us,
let us consider a simple example using quantification. Consider
the formula ∃x HAPPY(x ). This is meant to represent the circumstance that some individual is happy. In order for this to be true
in a model M with domain D and interpretation function I , there
would have to be some element of D that is happy according to I .
This means that the value I assigns to the constant HAPPY is the
characteristic function of a set that contains at least one element
of D. If I (HAPPY) is the characteristic function of the empty set,
then nobody is happy. But if I (HAPPY) is a function that returns
“true” for some individual, then somebody is happy.
For example, recall from (9) above, repeated in (14), that in
M 1 , there are two happy individuals, Bart and Maggie:
(14)
⎡ Bart
→ 1 ⎤⎥
⎢
⎢
⎥
I 1 (HAPPY) = ⎢ Homer → 0 ⎥
⎢
⎥
⎢ Maggie → 1 ⎥
⎣
⎦
So ∃x HAPPY(x ) ought to turn out to be true in M 1 .
The rule for existential quantification (rule 5(b) in (13)) ensures that ∃x HAPPY(x ) is true only when I (HAPPY) is a function
that returns “true” for some individual, because it asks us to find
(at least) one individual to assign x to such that HAPPY(x ) is true.
The rule specifies the semantics for an arbitrary formula of the
form ∃vφ, so to see this, we have to think about how it applies to
our particular case. Here, our v is x and our φ is HAPPY(x ). Plugging those values into the definition, we get that:
90
(15)
Logic
⟦∃x
HAPPY (x )⟧
M ,g
= 1 iff there is an individual k ∈ D such
= 1.
that: ⟦HAPPY(x )⟧
M ,g [x ↦k ]
What this says is that given a model M and an assignment function g , the sentence ∃x HAPPY(x ) is true with respect to M and
g if we can modify the assignment function g in such a way that
x has a denotation that makes HAPPY(x ) true. This is a general
statement about any arbitrary model M and assignment function
g . To make things more concrete, let’s take a particular model and
a particular assignment function. Suppose M is M 1 from above
and g is g 1 . To show that ∃x HAPPY(x ) is true with respect to M 1
and g 1 , we have to find a way of modifying g 1 in a way that makes
HAPPY (x ) true with respect to M 1 and the new assignment function.
The formula HAPPY(x ) has no quantifiers in it, and the variable x, in particular, is not bound by any quantifier. This makes x
a free variable. A formula with free variables is called an open formula. (A formula with no free variables is a closed formula.) In
order to give a truth value to an open formula, we need to assign
values to the free variables in it. Otherwise we can’t determine
whether it’s true or false. In particular, whether HAPPY(x ) is true
or not depends on how x is interpreted, and this is given by the
assignment function.
Given an assignment function which maps x to Bart, it should
turn out that HAPPY(x ) is true with respect to M 1 , because Bart is
happy in M 1 . So let’s choose a variant of g 1 where x is mapped to
Bart.
⎡
⎢
⎢
⎢
g 1 [x ↦ Bart] = ⎢
⎢
⎢
⎢
⎣
x → Bart ⎤
⎥
y → Bart ⎥
⎥
⎥
z → Bart ⎥
⎥
⎥
...
⎦
Let’s call this assignment function g Bart for short. Since this is an
assignment that maps x to Bart, it should be the case that HAPPY(x )
Logic
91
is true in M 1 using this assignment function, i.e., that
⟦HAPPY(x )⟧M1 ,g Bart = 1.
We can prove this using the rules in (13). The formula to be evaluated is of the form π(α), where π is a predicate and α is a term,
so we can use the Predication rule (rule 2 in (13)). By this rule, we
have that:
⟦HAPPY(x )⟧M1 ,g Bart = ⟦HAPPY⟧M1 ,g Bart (⟦x ⟧M1 ,g Bart )
By the rule for variables 1(b) in (13), we have that:
⟦x ⟧M1 ,g Bart = g Bart (x )
Since g Bart (x ) = Bart, this means that:
⟦x ⟧M1 ,g Bart = Bart.
By the rule for non-logical constants 1(a) in (13), we have that:
⟦HAPPY⟧M1 ,g Bart = I 1 (HAPPY)
M ,g
M ,g
Since ⟦HAPPY(x )⟧ 1 Bart is equal to ⟦HAPPY⟧ 1 Bart (⟦x ⟧
⟦HAPPY⟧M1 ,g Bart = I 1 (HAPPY) and ⟦x ⟧M1 ,g Bart = Bart,
M 1 ,g Bart
) and
⟦HAPPY(x )⟧M1 ,g Bart = I 1 (HAPPY)(Bart).
I 1 (HAPPY) was defined above as a function that maps Bart and
Maggie to 1, and Homer to 0. Therefore, I 1 (HAPPY)(Bart) = 1, and
M ,g
from this it follows that ⟦HAPPY(x )⟧ 1 Bart = 1.
Now let us return to the existentially quantified formula we
started out with, ∃x HAPPY(x ). According to the rule for existenM ,g
tial quantification in (13), ⟦∃x HAPPY(x )⟧ 1 1 = 1 iff there is an
M ,g [x ↦k ]
individual k ∈ D such that that ⟦HAPPY(x )⟧ 1 1
= 1. We have
just seen that Bart is a k that makes HAPPY(x ) true. Therefore,
92
Logic
∃x HAPPY(x ) is true with respect to M1 and g 1 . We have proved
the following statement:
⟦∃x HAPPY(x )⟧M1 ,g 1 = 1
Notice that we could have started out with any other assignment
function and gotten the same result, because the answer only depends on what x is assigned to, and we can assign x to whatever
we want; we don’t have to stick with what g 1 gives us. So the value
of ∃x HAPPY(x ) does not depend on the assignment function. If
it’s true with respect to M 1 and g 1 , then it’s true with respect to M 1
and any other assignment function. In general, the assignment
function does not affect the truth value of closed formulas. It only
makes a difference for formulas that contain free variables.
If we wanted to show that the formula ∀x HAPPY(x ) was true,
we would have to consider assignments of x to every element of
the domain, not just one. (To show that it is false is easier; then
you just have to find one unhappy individual.) If HAPPY(x ) turns
out to be true no matter what the assignment function maps x to,
then ∀x HAPPY(x ) is true. Otherwise it is false.
Exercise 20. Is ∀x HAPPY(x ) true in M 1 ? Give a one-sentence explanation why or why not.
Exercise 21. Consider the following formulas.
Logic
93
(a)
HAPPY ( M ) ∧ HAPPY ( M )
(g) ∃x
(b)
HAPPY ( K )
(h) ∃x ∃z
(c)
HAPPY ( M , M )
(i) ∃x
(d) ¬¬HAPPY(N)
(e) ∀x
HAPPY (x )
(f) ∀x
HAPPY ( y )
LOVES (x, x )
LOVES (x, z )
(j) ∃∃x
(k) ∃x
LOVES (x, z )
LOVES (x, x )
HAPPY ( M )
Questions:
(a) Which of the above are well-formed formulas of L1 ?
(b) Of the ones that are well formed in L1 , which of the above formulas have free variables in them? (In other words, which of
them are open formulas?)
Exercise 22. Recall our fantasy model
happy:
⎡ Bart
⎢
⎢
I f (HAPPY) = ⎢ Homer
⎢
⎢ Maggie
⎣
(a) What is ⟦x ⟧
M f ,g Bart
→ 1 ⎤⎥
⎥
→ 1 ⎥
⎥
→ 1 ⎥⎦
? Apply the rule for variables 1(b) in (13).
M f ,g Bart
(b) What is ⟦HAPPY⟧
stants 1(a) in (13).
M f ,where everybody is
? Apply the rule for non-logical con-
(c) Which semantic interpretation rule in (13) do you need to use
in order to put the meanings of HAPPY and x together, and
compute the denotation of HAPPY(x )?
94
Logic
(d) Use the rule you identified in your answer to the previous
question to put together the meanings you computed in the
M ,g
previous steps, explaining carefully why ⟦HAPPY(x )⟧ f Bart =
1.
Now let us consider what happens with a binary predicate. Assume as above that everybody loves themselves and nobody else
in M 1 .
(16)
⎡ Maggie
⎡
⎢
⎢
⎢ Maggie → ⎢⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎢
⎣
⎢
⎡ Maggie
⎢
⎢
⎢
⎢
⎢
I 1 (LOVES) = ⎢ Homer → ⎢ Homer
⎢
⎢
⎢ Bart
⎢
⎣
⎢
⎢
⎡
Maggie
⎢
⎢
⎢
⎢
⎢ Bart
→ ⎢ Homer
⎢
⎢
⎢
⎢ Bart
⎣
⎣
→
→
→
→
→
→
→
→
→
1
0
0
0
1
0
0
0
1
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
Let us consider the formula ∃x LOVES(x )(M) with respect to
M 1 and g 1 . Is it true or false? We have to use the rule for existential quantification in order to figure it out. Plugging in x for v,
LOVES (x )( M ) for φ, M 1 for M , and g 1 for g , we get that
⟦∃x LOVES(x )(M)⟧M1 ,g 1 = 1
if and only if there is an individual k ∈ D such that
⟦LOVES(x )(M)⟧M1 ,g 1 [x ↦k ] = 1.
Let’s think strategically about what value to choose for k. It should
be someone whom, when fed as the first argument to the function
denoted by LOVES, yields a function that maps the denotation of
M to true. We stipulated above in (7) that I 1 ( M ) = Maggie. So we
Logic
95
should choose a k that loves Maggie. Since everybody loves themselves in M 1 , we can set k equal to Maggie and get the sentence
to turn out true. So let’s take a variant of g 1 where x is mapped to
Maggie:
⎡
⎢
⎢
⎢
g 1 [x ↦ Maggie] = ⎢
⎢
⎢
⎢
⎣
x → Maggie
y → Bart
z → Bart
...
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
This happens to be the same as g 1 , so we can just call it g 1 . Using this assignment function, the value of LOVES(x )(M) is true in
M 1 . To show this, we have to apply the Predication rule twice.
Recall the Predication rule: “If π is a unary binary predicate and
M ,g
α is a term, then ⟦π(α)⟧
= ⟦π⟧M ,g (⟦α⟧M ,g ). The expression
LOVES (x ) is a (complex) unary predicate, and M is a term, so the
predication rule tells us that:
⟦LOVES(x )(M)⟧M1 ,g 1 = ⟦LOVES(x )⟧M1 ,g 1 (⟦M⟧M1 ,g 1 )
M ,g
To calculate ⟦LOVES(x )⟧ 1 1 , we can use the Predication rule again,
because LOVES is a binary predicate and x is a term, thus:
⟦LOVES(x )⟧M1 ,g 1 = ⟦LOVES⟧M1 ,g 1 (⟦x ⟧M1 ,g 1 )
M ,g
We know that ⟦x ⟧ 1 1 = Maggie because g 1 (x ) = Maggie. AccordM ,g
ing to the rule for non-logical constants, ⟦LOVES⟧ 1 1 = I 1 (LOVES),
which we defined as the ‘self’ function. Applied to Maggie, this
gives a function that maps Maggie to 1 and all other individuals to
0.
⎡ Maggie → 1 ⎤
⎢
⎥
⎢
⎥
I 1 (LOVES)(Maggie) = ⎢ Homer → 0 ⎥
⎢
⎥
⎢ Bart
→ 0 ⎥⎦
⎣
M ,g
We can thus conclude that ⟦LOVES(x )(M)⟧ 1 1 = 1. This means
M ,g [x ↦k ]
that there is a k such that ⟦LOVES(x )(M)⟧ 1 1
= 1 (namely
M 1 ,g 1
Maggie), and therefore ∃x ⟦LOVES(x )(M)⟧
= 1.
96
Logic
Exercise 23. Let g be defined such that x ↦ c, y ↦ b, and z ↦ a.
(a) Using the semantic rules and definitions for M 1 above, calculate:
(a) ⟦x ⟧
M 1 ,g
(b) ⟦M⟧
M 1 ,g
(c) ⟦LOVES⟧
M 1 ,g
(d) ⟦LOVES(x )(M)⟧
M 1 ,g
(b) List all of the value assignments that are exactly like g except possibly for the individual assigned to x, and label them
g 1 ...g n .
(c) For each of those value assignments g i in the set {g 1 , ..., g n },
M ,g
calculate ⟦LOVES(x )(M)⟧ i .
M ,g
(d) ⟦∀x LOVES(x )(M)⟧
= 1 iff for every value assignment g ′
′
such that g is exactly like g except possibly for the individM ,g ′
ual assigned to x by g ′ , ⟦LOVES(x )(M)⟧
= 1. So, what is
M ,g
⟦∀x LOVES(x )(M)⟧ ?
Exercise 24. If a formula has free variables then it may well be true
with respect to some assignments and false with respect to others.
Give an example of two variable assignments g i and g j such that
⟦LOVES(x )(M)⟧M ,g i ≠ ⟦LOVES(x )(M)⟧M ,g j .
One important feature of the semantics for quantifiers and
variables in first-order logic using assignment functions is that it
scales up to formulas with multiple quantifiers. Recall the quantifier scope ambiguity in Every linguist admires a philosopher that
we discussed at the beginning of the section. That sentence was
Logic
97
said to have two readings, which can be represented as follows:
∀x [LINGUIST(x ) → ∃ y [PHILOSOPHER( y ) ∧ ADMIRES( y )(x )]]
∃ y [PHILOSOPHER( y ) ∧ ∀x [LINGUIST(x ) → ADMIRES( y )(x )]]
We will spare the reader a step-by-step computation of the semantic value for these sentences in a given model. We will just
point out that in order to verify the first kind of sentence, with a
universal quantifier outscope an existential quantifier, one would
consider modifications of the input assignment for every member of the domain, and within that, modifications of the modified assignment for some element of the domain making the existential statement true. To verify the second kind of sentence, one
would consider a single modification of the input assignment for
the outer quantifier (the existential quantifier), and then modifications of that modified assignment for every member of the domain in order to evaluate the embedded universal statement. This
procedure will work for indefinitely many quantifiers.
4.4 Typed lambda calculus
Now we will add one key symbol to our logic: λ ‘lambda’. This is
an extremely useful symbol in semantics. With this symbol, along
with variables that can range over objects of complex types, we
can represent a lot of useful ideas. For example, take the prefix
non-, as in non-smoker. A non-smoker is someone who is not in
the set of smokers. If SMOKES denotes a function of type ⟨e, t ⟩, the
characteristic function of the set of smokers, then the meaning of
non-smoker can be represented as follows:
λx [¬SMOKES(x )]
This expression denotes a function which takes an individual (x)
and returns 1 if and only if x is not a smoker. The meaning of
non- can be represented as a function that takes as its argument
98
Logic
a predicate and then returns a new predicate which holds of an
individual if the individual does not satisfy the input predicate.
We can represent this function as follows, where P is a variable
that ranges over functions from individuals to truth values:
λP λx [¬P (x )]
The denotation of this expression is a function which takes as input a function (namely P ) and returns a new function (namely
λx [¬P ]. The new function takes an individual (namely x) and returns true if and only if x does not satisfy P . If we apply this function to the predicate SMOKES, then we get the function denoted by
λx [¬SMOKES(x )]. This is the meaning of non-smoker, assuming
that smoker denotes SMOKES.
The λ symbol will also let us represent the idea that a natural language quantifier like every expresses a relation between two
sets X and Y , with the help of variables that range over functions.
Let X and Y be variables ranging over functions from individuals
to truth types. The meaning of every can be represented like this:
λX λY [∀x [ X (x ) → Y (x )]]
This expression denotes a function which takes a function from
individuals to truth values (call it X ), and then returns another
such function (call it Y ), and returns 1 (true) if and only if every X
is a Y .
The denotation of every cellist would be the result of applying
this function to whatever cellist denotes:
λX λY [∀x [ X (x ) → Y (x )]](CELLIST)
This expression will turn out to be equivalent to:
λY [∀x [CELLIST(x ) → Y (x )]]
Thus the meaning of every, applied to the meaning of cellist, is a
function that is still hungry for another unary predicate. If we feed
Logic
99
it SMOKES, then we get a sentence that denotes a truth value:
λY [∀x [CELLIST(x ) → Y (x )]](SMOKES)
≡ ∀x [CELLIST(x ) → SMOKES(x )]
A crucial point to notice here is that the argument of the function
is itself a function. Functions that take other functions as arguments are called higher-order. So what we are dealing with here is
no longer first-order logic, but higher-order logic. Pretty fancy.
We can achieve this kind of generality by expanding the set of
syntactic categories of our language. Before, we had a finite inventory of syntactic categories: terms, unary predicates, binary
predicates, and formulas. We will now have an infinite set of syntactic categories, one for each type. Recall the set of types:
• e is a type
• t is a type
• If a and b are types, then ⟨a, b ⟩ is a type.
• Nothing else is a type.
Each of these infinitely-many types now corresponds to a syntactic category. And there will be a close match between the syntactic
category of an expression as a type and the kind of denotation it
has. An expression of type e will denote an individual, an expression of type t will denote a truth value, an expression of type ⟨e, t ⟩
will denote a function individuals to truth values, etc.
The syntax rule introducing λ-expressions into the language
is defined in terms of types:
(17)
Lambda abstraction (syntax)
If α is an expression of type a and u is a variable of type b
then λu [α] is an expression of type ⟨b, a ⟩.
100
Logic
For example, assuming that x is a variable of type e (since it denotes an individual, given an assignment), and SMOKES(x ) is an
expression of type t (since it denotes a truth value, given a model
and an assignment), it follows from (17) that λx [SMOKES(x )] is an
expression of type ⟨e, t ⟩, the type of functions from individuals to
truth values.
What λx [SMOKES(x )] denotes is a function that takes as input an individual (call it x), and yields ‘true’ if x smokes, and false
otherwise. In general, the semantics of λ-expressions is defined
as follows:6
(18)
Lambda abstraction (semantics)
If α is an expression of type a and u is a variable of type b
M ,g
then ⟦λu [α]⟧
is that function h from D b into D a such
M ,g [u →k ]
that for all objects k in D b , h (k ) = ⟦α⟧
(Recall that g [u ↦ k ] is the assignment that is just like g save that
the variable u gets mapped to the object k.)
Let us use this definition to calculate the denotation of the expression λx [SMOKES(x )]. Preview: This expression will turn out
to have exactly the same denotation as the expression SMOKES: a
function of type ⟨e, t ⟩ which returns 1 if and only if the input is a
smoker, i.e., for all M , g :
⟦λx [SMOKES(x )]⟧M ,g = ⟦SMOKES⟧M ,g
To see this, let us look carefully at (18), and see how α, a, u, and b
should be instantiated. Our α is SMOKES(x ), and that is an expression of type t , so a is instantiated as t in this case. Our u
is x, and that is a variable of type e, so our b is instantiated as
M ,g
e. So according to (18), ⟦λx [SMOKES(x )]⟧
is that function h
from D e into D t such that for all objects k in D e , h (k ) is equal
M ,g [x →k ]
to ⟦SMOKES(x )⟧
. Since the variable x is assigned to the
individual k:
⟦SMOKES(x )⟧M ,g [x →k ] = ⟦SMOKES⟧M ,g [x →k ] (k ) = I (SMOKES)(k )
6
We are following the definition given by Dowty et al. (1981), p. 100.
Logic
101
So for all k in D e , h (k ) = I (SMOKES)(k ). We know already that
⟦SMOKES⟧M ,g = I (SMOKES); in other words, SMOKES also denotes
that function h such that for all k in D e , h (k ) = I (SMOKES)(k ). So
λx [SMOKES(x )] and SMOKES are equivalent.7
Since λx [SMOKES(x )] is an expression of type ⟨e, t ⟩, it can appear anywhere that other expressions of type ⟨e, t ⟩ can appear. So,
both of the following are well-formed expressions (of type t ):
• λx [SMOKES(x )](B)
•
SMOKES ( B )
Furthermore, they are equivalent. The latter is true if and only
if the denotation of B is in the set characterized by SMOKES. The
former is true under exactly the same circumstances. In general,
an expression of the form λu [φ](α), where u is any variable and
α is an expression of the same type, is equivalent to a version of
φ with all instances of u replaced by α. This is the principle of of
β-reduction (‘beta-reduction’), also known as λ-conversion:
(19)
β-reduction principle
λu [φ](α) is equivalent to φ[u /α]
where φ[u /α] is the result of replacing all instances of u in
φ with α.
Because of this fact, we can reduce λx [SMOKES(x )](a ) to form
an equivalent expression SMOKES(a ). When we say that these expressions are “equivalent”, we mean that they have the same value
in every model. They are not exactly the same string of symbols,
of course, so it is a slight abuse of notation to write:
λx [SMOKES(x )](a ) = SMOKES(a )
It is more correct to write:
λx [SMOKES(x )](a ) ≡ SMOKES(a )
7
This is an instance of the principle of η-equivalence ‘eta-equivalnce’): λu [α]
is equivalent to α when u is not free in α and α is of the appropriate type.
102
Logic
where ≡ means ‘is equivalent to’.
Note that β-reduction is not an additional stipulation. It’s just
a fact that follows from the definitions we have already given.
There is one caveat to the rule of β-reduction. It only works if
α does not contain any free variables that occur in φ. For example,
consider the following expression:
λx [∀ y [LOVES(x, x ) → LOVES(x, y )]]( y )
If we take away λx and substitute y for x then we get:
∀ y [LOVES( y, y ) → LOVES( y, y )]
Through overly enthusiastic use of β-reduction, the variable y accidentally became bound by the universal quantifier ∀. Before we
can apply β-reduction in this case, we must change the name of
the variable bound by the universal quantifier thus:
λx [∀z [LOVES(x, x ) → LOVES(x, z )]]( y )
We changed y to z in the scope of the universal quantifier. This is
harmless, because a bound variable can be replaced by any other
bound variable without a change in meaning. This is the principle
of α-equivalence. Now when we apply β-reduction, we get the
following result:
∀z [LOVES( y, y ) → LOVES( y, z )]
where the variable y no longer clashes with the variable bound by
the universal quantifier.
4.4.1 Syntax of L λ
Let us now summarize our new logic, L λ , which is a version of the
typed lambda calculus developed by logicial Alonzo Church. The
types are defined recursively as follows: e and t are both types. If
a and b are types, then ⟨a, b ⟩ is a type; nothing else is a type. A
formula is an expression of type t .
Logic
(20)
103
Syntax of L λ
The set of expressions of type a, for any type a, is defined
recursively as follows:
1. Basic expressions (now more general!)
For each type a,
(a) the non-logical constants of type a are they symbols of the form c n,a for each natural number n.
(b) the variables of type a are the symbols of the form
v n,a for each natural number n.
2. Predication (now more general!)
For any types a and b, if α is an expression of type
⟨a, b ⟩ and β is an expression of type a then α(β) is an
expression of type b.
3. Negation
If φ is a formula, then so is ¬φ.
4. Binary Connectives
If φ and ψ are formulas, then so are ¬φ, [φ ∧ ψ], [φ ∨
ψ], [φ → ψ], and [φ ↔ ψ].
5. Quantification
If φ is a formula and u is a variable of any type, then
∀uφ and ∃uφ are formulas.
6. Lambda abstraction (new!)
If α is an expression of type a and u is a variable of
type b then λu [α] is an expression of type ⟨b, a ⟩.
4.4.2 Semantics
As in L1 , the semantic values of expressions in Lλ depend on a
model and an assignment function. As in L1 , a model M = ⟨D, I ⟩
is a pair consisting of the domain of individuals D and an interpretation function I , which assigns semantic values to each of the
non-logical constants in the language. For every type a, I assigns
an object of type a to every non-logical constant of type a.
104
Logic
Types are associated with domains. The domain of individuals
D e = D is the set of individuals, the set of potential denotations for
an expression of type e. The domain of truth values D t contains
just two elements: 1 ‘true’ and 0 ‘false’. For any types a and b,
D ⟨a,b ⟩ is the domain of functions from D a to D b .
Assignments provide values for variables of all types, not just
those of type e. An assignment thus is a function assigning to each
variable of type a a denotation from the set D a .
The semantic value of an expression is defined as follows:
(21)
Semantics of L λ
1. Basic expressions
(a) If α is a non-logical constant, then ⟦α⟧
(b) If α is a variable, then ⟦α⟧
M ,g
M ,g
= I (α).
= g (α).
2. Predication
If α is an expression of type ⟨a, b ⟩, and β is an expression of type a, then ⟦α(β)⟧ = ⟦α⟧(⟦β⟧).
3. Negation
M ,g
If φ is a formula, then ⟦¬φ⟧
= 1 iff ⟦φ⟧M ,g = 0.
4. Binary Connectives
If φ and ψ are formulas, then:
= 1 iff ⟦φ⟧M ,g = 1 and ⟦ψ⟧M ,g = 1.
M ,g
(b) ⟦φ ∨ ψ⟧
= 1 iff ⟦φ⟧M ,g = 1 or ⟦ψ⟧M ,g = 1.
M ,g
(c) ⟦φ → ψ⟧
= 1 iff ⟦φ⟧M ,g = 0 and ⟦ψ⟧M ,g = 1.
M ,g
(d) ⟦φ ↔ ψ⟧
= 1 iff ⟦φ⟧M ,g = ⟦ψ⟧M ,g .
(a) ⟦φ ∧ ψ⟧
M ,g
5. Quantification
(a) If φ is a formula and v is a varaible of type a then
⟦∀vφ⟧M ,g = 1 iff for all k ∈ D a :
⟦φ⟧M ,g [u ↦k ] = 1
Logic
105
(b) If φ is a formula and v is a variable of type a then
⟦∃vφ⟧M ,g = 1 iff there is an individual k ∈ D a such
that that:
⟦φ⟧M ,g [v ↦k ] = 1.
6. Lambda Abstraction
If α is an expression of type a and u a variable of type
M ,g
b then ⟦λu [α]⟧
is that function h from D b into D a
M ,g [u →k ]
such that for all objects k in D b , h (k ) = ⟦α⟧
.
Exercise 25. Consider the following expressions.
(a) λx [P (x )](a )
(b) λx [P (x )(a )]
(c) λx [R (a )( y )]
(d) λx [R (a )( y )](b )
(e) λx [R (a )(x )](b )
(f) λx [λy [R ( y )(x )]](b )
(g) λx [λy [R ( y )(x )]](b )(a )
(h) λx [λy [R ( y )(x )](b )](a )
(i) λX [∃x [P (x ) ∧ X (x )]](λy [R (a )( y )])
(j) λX [∃x [P (x ) ∧ X (x )]](λx [R (a )(x )])
(k) λX [∃x [P (x ) ∧ X (x )]](λy [R (x )( y )])
(l) λX [∃x [P (x ) ∧ X (x )]](Q )
(m) λX [∃x [P (x ) ∧ X (x )]]( X )
(n) λX [∃x [P (x ) ∧ X (x )](λx [Q (x )])]
106
Logic
For each of the above, answer the following questions:
(a) Is it a well-formed formula of L λ ?
(b) If yes, what is its type?
(c) If the formula is well-formed, give a completely β-reduced
(λ-converted) expression which is equivalent to it. Use αequivalence (relettering of bound variables) if necessary to
avoid variable clash.
Assume the following abbreviations:
• x is v 0,e
• y is v 1,e
• P is v 0,⟨e,t ⟩ , Q is v 1,⟨e,t ⟩ , and X is v 2,⟨e,t ⟩
• R is v 0,⟨e,⟨e,t ⟩⟩
•
N
is c 0,e and D is c 1,e
You can check your answers using the Penn Lambda Calculator
(but note that you have to write R (x )( y ) as R ( y, x ) there).
Exercise 26. Do example1.txt in the Penn Lambda Calculator.
Exercise 27.
(a) Partially define a model for L λ giving values (denotations) to
the constants LOVES, N, and D.
(b) Show that λx [LOVES(N)(x )](D) and its β-reduced version
Logic
107
R (N)(D) have the same semantic value in your model using
the semantic rules for L λ .
5
∣
Translating to lambda calculus
Lambda calculus provides a nice clean way of representing the
meanings of English words and phrases. In this chapter we will
pursue a slightly new strategy for assigning truth conditions to
natural language expressions: Rather than specifying the denotations of English words directly with expressions like:
⎡ Bart
→ 1 ⎤⎥
⎢
⎥
⎢
⟦snores⟧M1 = ⎢ Maggie → 1 ⎥
⎥
⎢
⎢ Homer → 0 ⎥
⎦
⎣
we will translate expressions of English into expressions of lambda
calculus, with statements like this:
snores ↝ λx [SNORES(x )]
And then we will let the semantics of the English expressions be
inherited from the semantics of lambda calculus. A given sentence will be true with respect to a model and an assignment function if its translation is true with respect to that model and assignment function.
This is like what Montague (1974b) did in his famous work entitled The Proper Treatment of Quantification in Ordinary English
(‘PTQ’ for short). There, unlike in English as a Formal Language
(Montague, 1974a), he specified a set of principles for translating
English into his own version of lambda calculus called Intensional
Logic. He was very clear that this procedure was only meant to be
109
110
Translating to lambda calculus
a convenience; one could in principle specify the meanings of the
English expressions directly. So we will continue to think of our
English expressions as having denotations, even though we will
specify them indirectly via a translation to lambda calculus. Nevertheless, the expressions of lambda calculus are not themselves
the denotations. Rather, we have two object languages, English (a
natural language) and lambda calculus (a formal language), and
we are translating from the natural language to the formal language, and specifying the semantics of the formal language in our
metalanguage (which is also English, mixed with talk of sets and
relations).1
In this section, we will explore the limits of Functional Application, and define a larger fragment. To do so, we will need a few
more syntax rules.
(1)
Syntax
S
S
JP
S
VP
AP
NP
N′
N′
PP
→
→
→
→
→
→
→
→
→
→
NP VP
S JP
JS
Neg S
V (NP∣AP∣PP)
A (PP)
D N′
N (PP)
A N′
P NP
The vertical bar ∣ separates alternative possibilities, and the parentheses signify optionality, so the VP rule means that a VP can consist solely of a verb, or of a verb followed by an NP, or of a verb
1
An important difference between the tact we are taking here and the one
taken in Heim & Kratzer’s (1998) textbook is that here the λ symbol is used as an
expression of lambda calculus (its original use), whereas in Heim and Kratzer the
λ symbol is part of the meta-language, as an abbreviation for describing functions. One should carefully distinguish between these two ways of using it.
Translating to lambda calculus
111
followed by an AP, etc.
The terminal nodes of the syntax trees produced by these syntax rules may be populated by the following words:
(2)
Lexicon
J: and, or
Neg: it is not the case that
V: smokes, loves, is
A: lazy, proud
N: drunkard, baby, kid, zebra
D: the, a, every, some, no
P: of, with
NP: Bart, Maggie, Homer, everybody, somebody, nobody
Exercise 1. Which of the following strings are sentences of the
fragment of English that we have defined (modulo sentenceinititial capitalization)?
(a) Homer Simpson loves everybody.
(b) Some drunkard smokes every lazy zebra.
(c) Maggie is not a zebra.
(d) Homer Simpson is.
(e) Homer is.
(f) Homer is a baby zebra.
(g) Somebody is proud of the kid.
(h) A zebra loves proud of Bart.
(i) The proud zebra of Homer loves every lazy lazy lazy lazy
drunkard.
112
Translating to lambda calculus
(j) Bart smokes with nobody.
(k) Maggie and Bart are with Homer.
5.1 Fun with Functional Application
5.1.1 Homer is lazy
Let us consider how to analyze a sentence with an adjective following is, such as Homer is lazy. The syntactic structure is as follows:
S
NP
Homer
VP
V
AP
is lazy
Let us assume that the proper name Homer is translated as a
constant of type e; call it H.
(3)
Homer ↝ H
We can assume that lazy denotes a function of type ⟨e, t ⟩, the characteristic function of a set of individuals (those that are lazy).Let
us use LAZY as an abbreviation for that constant of type ⟨e, t ⟩. This
will be the translation for lazy.
(4)
lazy ↝ LAZY
What is the meaning of is? It does not seem to contribute
much information in this sentence. We can make do by treating it
as an identity function, a function that returns whatever it takes
in as input. In this case, is takes in a function of type ⟨e, t ⟩, and
returns that same function.
Translating to lambda calculus
(5)
113
is↝ λP [P ]
This implies that is denotes a function that takes as its first argument another function P , where P is of type ⟨e, t ⟩, and returns P .
The next question is how to treat the constituent is lazy. Since
we are translating expressions of English into lambda calculus, we
have to specify our semantic composition rules in terms of translations now. We will associate translations not only with words,
but also with syntactic trees. We can think of words as degenerate
cases of trees, so in general, translations go from trees to expressions of our logic.
Composition Rule 1. Functional Application (FA)
If γ is a tree whose only two subtrees are α and β, and α ↝ α′ ,
where α′ is type ⟨a, b ⟩ and β ↝ β′ , where β′ is of type a, then:
γ ↝ α′ (β′ ).
For example, the tree [VP [V is ] [AP lazy ] ] has two subtrees, [V is
] and [AP lazy ]. If the translation of first one denotes a function
that applies to the other, then we can use FA to put together the
meanings of these two expressions. In order to give these two a
translation, we will also need a rule for non-branching nodes.
Composition Rule 2. Non-branching Nodes (NN)
If α ↝ α′ , and β is a tree whose only daughter is α, then β ↝ α′ .
This tells us that, for example, is has the same translation as [ V
is ], and that lazy has the same translation as [AP lazy ]. For thoroughness, we also define a rule covering terminal nodes (nodes
in the tree that have no daughters, i.e., the words). Later on, we
will make a distinction between lexical terminals (content words)
114
Translating to lambda calculus
and non-lexical terminals (pronouns and traces), and the rule we
are introducing here is only appropriate for the former, hence the
name of the rule.
Composition Rule 3. Lexical Terminals (LT)
If α is a terminal node, then its translation is specified in its lexical
entry.
With these rules, we will end up with the following analysis for the
sentence Homer is lazy:
S: t
(6)
LAZY ( H )
NP: e
VP: ⟨e, t ⟩
H
LAZY
Homer
V: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
λP [P ]
A: ⟨e, t ⟩
is
lazy
LAZY
Each node shows the syntactic category, the semantic type, and a
fully β-reduced translation to lambda calculus. In this case, Functional Application is used at all of the branching nodes (S and VP),
Non-branching Nodes is used at all of the non-branching nonterminal nodes (NP, V, and A), and Lexical Terminals is used at all
of the terminal nodes (Homer, is, and lazy).
5.1.2 Homer loves Maggie
We can of course use Functional Application to analyze transitive
verbs as well. Let us use the constant LOVES as the translation of
the verb loves. Let us assume further that Homer is translated as D
and that Maggie is translated as M.
Translating to lambda calculus
115
• loves ↝ LOVES
• Homer ↝ D
• Maggie ↝ M
Homer loves Maggie can then be analyzed as follows.
S: t
(7)
VP: ⟨e, t ⟩
NP: e
Homer
V: ⟨e, ⟨e, t ⟩⟩
NP: e
loves
Maggie
Via Functional Application, the transitive verb loves combines with
the object Maggie, and the VP combines with the subject Homer.
The translation of the VP is an expression of type ⟨e, t ⟩, denoting a
function from individuals to truth values. This applies to the denotation of Homer (an individual) to produce a truth value.
Exercise 2. Give translations for each node of the tree in (7).
5.1.3 Homer is with Maggie
Like adjectives, prepositional phrases can also serve as predicates,
as in, for example, Homer is with Maggie. Let us assume that with
translates as the constant WITH, a function of type ⟨e, ⟨e, t ⟩⟩.
Here is an overview of how the derivation will go. Via Functional Application, the preposition with combines with its object
Maggie, and the resulting PP combines with is to form a VP which
combines with tehe subject Homer. The translation of the VP is an
expression of type ⟨e, t ⟩, denoting a function from individuals to
truth values. This applies to the denotation of Homer to produce
a truth value.
116
Translating to lambda calculus
S: e
NP: e
VP: ⟨e, t ⟩
Homer
V: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
is
PP: ⟨e, t ⟩
P: ⟨e, ⟨e, t ⟩⟩
NP: e
with
Maggie
Exercise 3. Derive the translation into lambda calculus for the full
sentence by giving a fully β-reduced translation for each node.
Exercise 4. Suppose model M 1 is defined such that the following
holds:
⎡ Bart
⎡
⎢
⎢
⎢
⎢ Bart
→ ⎢ Maggie
⎢
⎢
⎢
⎢ Homer
⎢
⎣
⎢
⎢
⎡ Bart
⎢
⎢
⎢
M ⎢
⟦WITH⟧ = ⎢ Maggie → ⎢ Maggie
⎢
⎢
⎢
⎢ Homer
⎢
⎣
⎢
⎡ Bart
⎢
⎢
⎢
⎢ Homer → ⎢⎢ Maggie
⎢
⎢
⎢
⎢ Homer
⎣
⎣
M
M
→
→
→
→
→
→
→
→
→
0
0
1
0
0
0
1
0
0
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎦
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
and suppose that ⟦D⟧ 1 =Homer and ⟦M⟧ 1 =Maggie. What truth
value does the translation for Homer is with Maggie have in M 1 ?
Explain your answer.
Translating to lambda calculus
117
5.1.4 Homer is proud of Maggie
Like prepositions, adjectives can denote functions of type ⟨e, ⟨e, t ⟩⟩.
Proud is an example; in Homer is proud of Maggie, the adjective
proud expresses a relation that holds between Homer and Maggie. We can capture this by assuming that proud translates to the
constant PROUD of type ⟨e, ⟨e, t ⟩⟩, denoting a function that takes
two arguments, first a potential object of pride (such as Maggie),
then a potential bearer of such pride (e.g. Homer), and then returns if the pride relation holds between them.
In contrast to with, the preposition of does not seem to contribute much content. We can assume that of is a function word
like is, and also denotes an identity function. Unlike is, however,
we will treat of as an identity function that takes an individual and
returns an individual, so it will be of category ⟨e, e ⟩.
(8)
of ↝ λx [x ]
So the adjective phrase proud of Maggie will have the following
structure:
AP: ⟨e, ⟨e, t ⟩⟩
A: ⟨e, ⟨e, t ⟩⟩
proud
PP: e
P: ⟨e, e ⟩
NP: e
of
Maggie
Exercise 5. Give a fully β-reduced form of the translation at each
node for Homer is proud of Maggie.
Exercise 6. What would go wrong if we were to assume that of was
of type ⟨e, ⟨e, t ⟩⟩, like with?
118
Translating to lambda calculus
5.1.5 Homer is a drunkard
Let us consider Homer is a drunkard. The noun drunkard can
be analyzed as an ⟨e, t ⟩ type property like lazy, the characteristic
function of the set of individuals who are drunkards.
The indefinite article a is another function word that does not
seem to be contributing much content. We will assume that a,
like is, denotes a function that takes an ⟨e, t ⟩-type predicate and
returns it.
(9)
a ↝ λP [P ]
With these assumptions, the derivation will go as follows.
S: t
(10)
VP: ⟨e, t ⟩
NP: e
Homer
V: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
is
NP: ⟨e, t ⟩
D: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
N′ : ⟨e, t ⟩
a
drunkard
Exercise 7. Give fully β-reduced translations at each node of the
tree for Homer is a drunkard.
Exercise 8. Can we treat a as ⟨⟨e, t ⟩, ⟨e, t ⟩⟩ in a sentence like A
drunkard loves Maggie? Why or why not?
Translating to lambda calculus
119
5.2 Predicate Modification
5.2.1 Homer is a lazy drunkard
Now let us consider Homer is a lazy drunkard. We have a little
bit of a problem on our hands right now. According to our lexical
entry for lazy above, lazy denotes a function of type ⟨e, t ⟩. How
can we combine that with something of type ⟨e, t ⟩?
(11)
S: t
VP: ⟨e, t ⟩
NP: e
Homer
NP: ⟨e, t ⟩
V: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
is
N′ : ⟨e, t ⟩
D: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
a
A: ⟨e, t ⟩
N′ : ⟨e, t ⟩
lazy
drunkard
If we want to use Functional Application here, we need lazy to be a
function of type ⟨⟨e, t ⟩, ⟨e, t ⟩⟩. We can do this using a type-shifting
rule which introduces a possible translation of type ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
for every translation of type ⟨e, t ⟩ (Partee, 1995, p. 29).
Type-Shifting Rule 1. Predicate-to-modifier shift
If α ↝ α′ , where α′ is of category ⟨e, t ⟩, then:
α ↝ λP λx [α′ (x ) ∧ P (x )]
as well (as long as P and x are not free in α′ ; in that case, use different variables of the same type).
120
Translating to lambda calculus
By this rule, we can derive that:
lazy ↝ λP λx [LAZY(x ) ∧ P (x )].
so Homer is a lazy drunkard gets the following analysis:
(12)
S: t
LAZY ( H ) ∧ DRUNKARD ( H )
NP: e
VP: ⟨e, t ⟩
λx [LAZY(x ) ∧ DRUNKARD(x )]
H
Homer
NP: ⟨e, t ⟩
λx [LAZY(x ) ∧ DRUNKARD(x )]
V: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
λP [P ]
is
D: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
λP [P ]
N′ : ⟨e, t ⟩
λx [LAZY(x ) ∧ DRUNKARD(x )]
a
N′ : ⟨e, t ⟩
A: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
λP λx [LAZY(x ) ∧ P (x )]
DRUNKARD
⇑
⟨e, t ⟩
drunkard
LAZY
lazy
This is essentially Montague’s strategy for dealing with adjectives
that modify nouns. In fact, all adjectives are ⟨⟨e, t ⟩, ⟨e, t ⟩⟩ for him.
For a sentence like Homer is lazy, Montague assumes that the adjective combines with a silent noun. We would have to augment
our syntax rules to allow for this, but if we did, then it would become mysterious why *Homer is lazy drunkard is not grammatical
in English, and why we can’t have silent nouns all over the place.
So it seems necessary to assume that adjectives can at least sometimes be of type ⟨e, t ⟩.
Translating to lambda calculus
121
Instead of using a type-shifting rule to interpret attributive adjectives, another strategy is to let go of Frege’s conjecture (that
Functional Application is basically the only composition rule), and
accept another composition rule. This rule would take two predicates of type ⟨e, t ⟩, and combine them into a new predicate of
type ⟨e, t ⟩. The new predicate would hold of anything that satisfied both of the old predicates. This is how Predicate Modification
is defined:
Composition Rule 4. Predicate Modification (PM)
If α ↝ α′ and β ↝ β′ and α′ and β′ are of type ⟨e, t ⟩, and γ is a tree
whose only two subtrees are α and β, then:
γ ↝ λu [α′ (u ) ∧ β′ (u )]
where u is a variable of type e that does not occur free in α′ or β′ .
This gives us the following derivation for the N′ :
(13)
N′ : ⟨e, t ⟩
λx [LAZY(x ) ∧ DRUNKARD(x )]
A: ⟨e, t ⟩
N′ : ⟨e, t ⟩
LAZY
DRUNKARD
lazy
drunkard
Exercise 9. Consider the sentence Maggie is a lazy baby. Give
two different analyses of the sentence, one using the Predicateto-modifier shift, and one using Predicate Modification. Give your
analysis in the form of a tree that shows for each node, the syntactic category, the type, and a fully β-reduced translation. (Feel free
to use the Lambda Calculator for this.)
122
Translating to lambda calculus
Exercise 10. Homer is a former drunkard does not entail Homer
is a drunkard and *Homer is former. In this sense, former is
a non-intersective modifier. Which of the following are nonintersective modifiers? Give examples to support your point.
(a) yellow
(b) round
(c) alleged
(d) future
(e) intelligent
(f) good
5.3 Quantifiers
5.3.1 Quantifiers: not type e
Previously, we analyzed indefinite descriptions like a baby in sentences like Maggie is a baby. But we still cannot account for the
use of a baby as the subject of a sentence or the object of a transitive verb, as in the following sentences:
(14)
a.
A baby loves Homer.
b.
Homer loves a baby.
We have analyzed the meaning of a as an identity function on
predicates, so a baby denotes a predicate (i.e., a function of type
⟨e, t ⟩). In a sentence like A baby loves Homer, the VP loves Homer
denotes a predicate of type ⟨e, t ⟩. This leaves us in the following
predicament.
Translating to lambda calculus
123
S: ?
VP: ⟨e, t ⟩
NP: ⟨e, t ⟩
D: ⟨⟨e, t ⟩, ⟨e, t ⟩⟩
N: ⟨e, t ⟩
V : ⟨e, ⟨e, t ⟩⟩
NP: e
a
baby
loves
Homer
The only composition rule that we have for combining two expressions of type ⟨e, t ⟩ is Predicate Modification. But this yields a
predicate, and it doesn’t make sense to analyze a sentence like A
baby loves Homer as denoting a predicate, because the sentence is
something that can be true or false. Its semantic value in a model
should be true or false, depending on whether or not a baby loves
Homer.
We also lack an analysis for the following sentences.
(15)
Somebody is lazy.
(16)
Everybody is lazy.
(17)
Nobody is lazy.
(18)
{Some, every, at least one, at most one, no} linguist is lazy.
(19)
{Few, some, several, many, most, more than two} linguists
are lazy.
We have been assuming that the VP denotes a function of type
⟨e, t ⟩. A sentence is something that can be true or false, so that
should be of type t . So what type is the subject in these examples?
S: t
NP: ?
VP: ⟨e, t ⟩
...
is lazy
124
Translating to lambda calculus
We could arrive at type t at the S node by treating these expressions as type e, like Homer, Maggie, and the baby. But there is
considerable evidence that these expressions cannot be treated as
type e. The analysis of these expressions as type e makes a number of predictions that are not borne out.
First, an individual-denoting should validate subset-to-superset
inferences, for example:
(20)
John came yesterday morning.
Therefore, John came yesterday.
This is a valid inference if the subject, like John, denotes an individual. Here is a proof. Everything that came yesterday morning
came yesterday. If α denotes an individual, then α came yesterday
morning is true if the individual denoted by α is among the things
that came yesterday morning. If that is true, then that individual
is among the things that came yesterday. So if the first sentence is
true, then the second sentence is true. QED.
Some of the expressions in (15)–(19) fail to validate subset-tosuperset inferences. For example:
(21)
At most one letter came yesterday morning.
Therefore, at most one letter came yesterday.
This inference is not valid, so at most one letter must not denote
an individual.
Exercise 11. Which of the expressions in (15)–(19) validate subsetto-superset inferences? Give examples.
The second property that expressions of type e should have
that these expressions do not always have is that they validate the
law of contradiction. In logic, the law of contradiction is that P ∧
¬P is always false. In this context, the prediction is that sentences
like the following should be self-contradictory:
Translating to lambda calculus
(22)
125
Mount Rainier is on this side of the border, and Mount Rainier
is on the other side of the border.
This sentence is contradictory because Mount Rainier denotes an
individual. Here is a proof. Nothing that is on this side of the border is on the other side of the border. If α is of type e, then α is
on this side of the border is true if and only if the individual that α
denotes is on this side of the border. This means that this individual is not on the other side of the border. So the second conjunct
must be false. So the conjunction (under a standard analysis of
and) can never be true. QED.
But the following sentence is not contradictory:
(23)
More than two mountains are on this side of the border,
and more than two mountains are on the other side of the
border.
So more than two mountains must not be type e.
Exercise 12. Which of the expressions in (15)–(19) fail to validate
the law of contradiction? Give examples.
Finally, an expression of type e should validate the law of the
excluded middle. The law of the excluded middle says that either
P is true, or ¬P is true. It can’t be the case that neither is true. For
example:
(24)
I am over 30 years old, or I am under 40 years old.
This is a tautology, and that is because I is an expression of type
e. Any expression of type e will yield a tautology in a sentence like
this. Here is a proof: Everything is either over 30 years old or under
40 years old. If α is of type e, then α is over 30 years old means that
the individual that α denotes is over 30 years old. α is under 40
years old means that the individual is under 40 years old. Since
everything satisfies at least one of these criteria, the disjunction
(under a standard analysis of or) cannot fail to be true.
126
Translating to lambda calculus
But this sentence is not a tautology:
(25)
Every woman in this room is over 30 years old, or every
woman in this room is under 40 years old.
So every woman must not be of type e.
Exercise 13. Which of the expressions in (15)–(19) fail to validate
the law of the excluded middle? Give examples.
5.3.2 Solution: Generalized quantifiers
Let us recap. We assume that a VP denotes a predicate (type ⟨e, t ⟩)
and that a sentence denotes a truth value (type t ). We have a
bunch of expressions that are not of type e, and we want them
to combine with the VP to produce something of type t . What can
we do? The solution is to feed the VP as an argument to the subject NP. An NP like something will denote a function that takes a
predicate as an argument, and returns a truth value. Its type will
therefore be:
⟨⟨e, t ⟩, t ⟩
This is the type of a quantifier.
We can define something as a function that takes as input a
predicate and returns true if and only if there is at least one satisfier of the predicate:
(26)
something ↝ λP [∃x [P (x )]]
In contrast, nothing returns true if there is nothing satisfying the
predicate:
(27)
nothing ↝ λP [¬∃x [P (x )]]
Everything returns true if everything satisfies the predicate:
(28)
everything ↝ λP [∀x [P (x )]]
Translating to lambda calculus
127
You can think of quantifiers as predicates of predicates. For example, λP [¬∃x [P (x )]] is a predicate that holds of a predicate P if P
has no satisfiers.
Using Functional Application (which, as you may recall, does
not care about the order of the arguments), the quantifier will take
the denotation of the VP as an argument, and yield a truth value,
thus:
S: t
S: t
vs.
DP: ⟨⟨e, t ⟩, t ⟩
VP: ⟨e, t ⟩
DP: e
VP: ⟨e, t ⟩
everything
V: ⟨e, t ⟩
Maggie
V: ⟨e, t ⟩
vanished
vanished
Now what about determiners like every, no, and some? These
should denote functions that take the denotation of a noun phrase
and return a quantifier, because we want every girl to function in
the same way as everyone. So the type of these kinds of determiners will be:
⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩
In particular, these determiners can be defined as follows:
(29)
some ↝ λP λQ [∃x [P (x ) ∧ Q (x )]]
(30)
no ↝ λP λQ [¬∃x [P (x ) ∧ Q (x )]]
(31)
every ↝ λP λQ [∀x [P (x ) → Q (x )]]
An expression of type ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩ is a generalized quantifier
(Barwise & Cooper, 1981). These will yield analyses like the following:
128
Translating to lambda calculus
S: t
∀x [DOUBT(x ) → VANISHED(x )]
VP: ⟨e, t ⟩
DP: ⟨⟨e, t ⟩, t ⟩
λQ [∀x [DOUBT(x ) → Q (x )]]
VANISHED
D: ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩
λP λQ [∀x [P (x ) → Q (x )]]
NP: ⟨e, t ⟩
VANISHED
every
doubt
V: ⟨e, t ⟩
DOUBT
vanished
Exercise 14. Give an analysis of A baby loves Homer using Functional Application, Non-Branching Nodes, and Lexical Terminals.
Your analysis should take the form of a tree, specifying at each
node, the syntactic category, the semantic type, and a fully βreduced translation to L λ . The sentence should be true in any
model where there is some individual that is both a baby and
someone who loves Homer. You may have to introduce a new lexical entry for a.
6
∣
Variables
6.1 Relative clauses
Now we will consider relative clauses such as the following:
(1)
The car that Joe bought is very fancy.
(2)
The woman who admires Joe is very lovely.
Semantically, relative clauses are much like intersective adjectives:
(3)
The red car is very fancy.
(4)
The Swedish woman is very lovely.
They are type ⟨e, t ⟩ and can combine via Predicate Modification.
NP: ⟨e, t ⟩
(5)
NP: ⟨e, t ⟩
CP: ⟨e, t ⟩
car
that Joe bought
CP stands for “Complementizer Phrase” and we will follow Heim &
Kratzer (1998) in assuming the following syntax for relative clause
CPs.
129
130
Variables
(6)
CP
C′
whichi
C
that
S
DP
VP
V
DP
bought
ti
Joe
(7)
CP
C′
whoi
C
that
S
VP
DP
ti
V
DP
likes
Joe
The text that is struck out like so is deleted. Following Heim &
Kratzer (1998), we assume that either the relative pronoun or the
complementizer that is deleted, in accordance with the ‘DoublyFilled Comp Filter’ (Chomsky & Lasnik, 1977), the principle that
either the relative pronoun or that must be silent in English.
These structures are assumed to be derived through a syntactic transformation which moves a relative pronoun from its original position to a position preceding the complementizer position,
C, where it is a sister to C′ .1 The thought is that which and who
1
This position is called the specifier position. The term ‘specifier’ comes from
the X -bar theory of syntax, where all phrases are of the form [ X P (specifier) [ X ′
[ X (complement) ] ] ] ].
Variables
131
occupy these original positions at Deep Structure, and then undergo a syntactic transformation to end up in their overt positions
at Surface Structure. But these wh- pronouns do not disappear
entirely from their original positions; they leave a trace signifying
that they once were there. The little subscript i on which and the
trace is there to keep track of the connection between the wh- pronoun and its original position. It will be instantiated as a natural
number (0,1,2,...) in any particular case, and this index plays an
important role in the interpretation.
Following Heim & Kratzer (1998), let us assume that traces are
interpreted as variables, so their denotation depends on an assignment. Here we implement this idea by mapping traces to variables in Lλ . The index on the trace determines which variable it is
translated to.
Composition Rule 5. Pronouns and Traces Rule
If α is an indexed trace or a pronoun, αi ↝ v i ,e
For the present purposes, we will assume that all pronouns and
traces are type e. (We could loosen this by adding a subscript for
the type on the pronoun or trace.) So for example,
t 5 ↝ v 5,e
I.e., t 5 is translated as variable number 5 of type e. The denotation
of this variable will depend of course on the assignment; recall
M ,g
that ⟦v 5,e ⟧
= g (v 5,e ).
We have called it the ‘Pronouns and Traces Rule’ because it
will also use it for pronouns, so for example:
he7 ↝ v 7,e
More on this in §6.3. Note that the idea of treating traces and
pronouns as variables is a rather controversial idea; see Jacobson
132
Variables
(1999) and Jacobson (2000) for critique and alternatives. But since
this is such an influential approach, it is important to understand
how it works.
Now, the S in a relative clause is type t . How do we get the CP
to have type ⟨e, t ⟩?
CP : ⟨e, t ⟩
(8)
C′ :
which1
S: t
C:
that
DP: e
Joe
VP: ⟨e, t ⟩
V: ⟨e, ⟨e, t ⟩⟩
DP: e
bought
t1
Heim & Kratzer (1998) propose the following, where p is a variable
of type t :
• The complementizer that is vacuous; that S = S
or that ↝ λp t [p ]
• The relative pronoun triggers a special rule called Predicate
Abstraction at the CP node.2
2
Heim and Kratzer’s original formulation of Predicate Abstraction is as follows: “If α is a branching node whose daughters are a relative pronoun indexed
x /i
i and β, then ⟦α⟧g = λ x ∈ De . ⟦β⟧g ,” where g x /i is an assignment that is just
like g except that x is assigned to i . In their framework, natural language expressions are mapped to denotations directly rather than being translated to lambda
calculus, and the mapping may depend on an assignment. The λ symbol is part
of the meta-language in their system, used as shorthand for describing functions, and x is also a variable of the meta-language, which ranges over objects in
the domain.
Variables
133
Composition Rule 6. Predicate Abstraction
If γ is an expression whose only two subtrees are αi and β and
β ↝ β′ where β′ is an expression of type t , then γ ↝ λv i ,e [β′ ].
(We will leave off e from the subscripts on variables when it is clear
what the type is.)
This gives us the following analysis of the relative clause:
(9)
CP : ⟨e, t ⟩
λv 1 [BOUGHT(v 1 )(J)]
which1
C′ : t
BOUGHT (v 1 )( J )
C: ⟨t , t ⟩
λp t [p ]
that
S: t
BOUGHT (v 1 )( J )
DP: e
VP: ⟨e, t ⟩
J
BOUGHT (v 1 )
Joe
V: ⟨e, ⟨e, t ⟩⟩
BOUGHT
DP: e
v1
bought
t1
So that Joe bought denotes the property of having bought by Joe.
This can combine via Predicate Modification with car, giving the
property of being a car that Joe bought, as it should.
6.2 Quantifier Raising
The rule of Predicate Abstraction will also help with the interpretation of quantifiers.
134
Variables
6.2.1 The problem of quantifiers in object position
Recall that a sentence like every linguist offended John, with a quantificational noun phrase in subject position, receives an analysis
like this:
S: t
(10)
VP: ⟨e, t ⟩
NP: ⟨⟨e, t ⟩, t ⟩
D: ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩
N′ : ⟨e, t ⟩
every
linguist
V: ⟨e, ⟨e, t ⟩⟩
NP: e
offended
John
But a problem arises when the quantifier is in object position. The
transitive verb is expecting an individual, so the quantifier phrase
cannot be fed as an argument to the verb. And the quantifier
phrase is expecting an ⟨e, t ⟩-type predicate, so the verb cannot be
fed as an argument to the quantifier phrase. So we have a type
mismatch.
(11)
S: ???
NP: e
VP: ???
John
V: ⟨e, ⟨e, t ⟩⟩
offended
NP: ⟨⟨e, t ⟩, t ⟩
D: ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩
N′ : ⟨e, t ⟩
every
linguist
There are several types of approaches to the problem:
1. Quantifier Raising: Assume a syntactic transformation such
that the quantifier phrase moves to position in the tree where
Variables
135
it can be interpreted, leaving a DP trace of type e in object
position.
2. Cooper Storage: Introduce a storage mechanism into the
semantics (Cooper, 1983). (This is done in Head-Driven Phrase
Structure Grammar; see Pollard & Sag 1994.) Heim & Kratzer
(1998) describe this as a way of “simulating movement” in
the semantics.
3. Flexible types approach: Interpret the quantifier phrase in
situ, i.e., in the position where it is pronounced. In this case
one can apply a type-shifting operation to change either the
type of the quantifier phrase or the type of the verb.
The flexible types approach adheres to the principle of “Direct
Compositionality”, which rejects the idea that the syntax first builds
syntactic structures which are then sent to the semantics for interpretation as a second step. With direct compositionality, the
syntax and the semantics work in tandem, so that the semantics
is computed as sentences are built up syntactically, as it were. Jacobson (2012) argues that this is a priori the simplest hypothesis
and defends it against putative empirical arguments against it.
Heim & Kratzer (1998) give two versions of a flexible types approach and give empirical arguments against both of them, in favor of the QR approach. Here we will present the QR approach
and compare it to flexible types approaches that Heim & Kratzer
(1998) do not consider, and argue that the available evidence does
not clearly favor or disfavor QR.
6.2.2 A Quantifier Raising approach
The Quantifier Raising solution to the problem of quantifiers in
object position is embedded in a syntactic theory with several levels of representation:
• Deep Structure (DS): Where the derivation begins, and active sentences (John kissed Mary) look the same as passive
136
Variables
sentences (Mary was kissed by John), and wh- words are in
their original positions. For example, Who did you see? is
You did see who? at Deep Structure.
• Surface Structure (SS): Where the order of the words corresponds to what we see or hear (after e.g. passivization or
wh-movement)
• Phonological Form (PF): Where the words are realized as
sounds (after e.g. deletion processes)
• Logical Form (LF): The input to semantic interpretation (after e.g. QR)
Transformations map from DS to SS, and from SS to PF and LF:
DS
SS
LF
PF
This is the so-called ‘T-model’, or ‘Y-model’. (Since the transformations from SS to LF happen “after” the order of the words is
determined, we do not see the output of these transformations.
These movement operations are in this sense covert.)
Under Heim & Kratzer’s (1998) assumptions, a transformation
called QR (Quantifier Raising) maps the SS structure in (12a) to
the LF structure in (12b).
(12)
a.
S
VP
DP
John
V
offended
DP
D
NP
every
linguist
Variables
137
b.
S
λP
DP
D
NP
every
linguist
1
S
VP
DP
John
V
DP
offended
t1
The number 1 in the syntax tree is the variable index that will be
used in the Predicate Abstraction rule. (Heim and Kratzer do not
give a syntactic category to the node dominating this index, so I
have called it λP.) The derivation works as follows. Predicate Abstraction is used at the λP node; Functional Application is used at
all other branching nodes.
(13)
S: t
∀x [LINGUIST(x ) → OFFENDED(x )(J)]
DP: ⟨⟨e, t ⟩, t ⟩
λQ [∀x [LINGUIST(x ) → Q (x )]]
D: ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩
λP λQ [∀x [P (x ) → Q (x )]]
NP: ⟨e, t ⟩
LINGUIST
every
linguist
λP: ⟨e, t ⟩
λv 1 [OFFENDED(v 1 )]
S: t
1
OFFENDED (v 1 )( J )
DP: e
VP: ⟨e, t ⟩
J
OFFENDED (v 1 )
John
V: ⟨e, ⟨e, t ⟩⟩
OFFENDED
DP: e
v1
offended
t1
138
Variables
Exercise 1. Simplify the following expression step-by-step:
λQ [∀x [LINGUIST(x ) → Q (x )]](λv 1 [OFFENDED(v 1 )])
Tip: Use the Penn lambda calculator.
Exercise 2. Derive a translation into lambda calculus for Beth
speaks a European language. Assume that a European language
undergoes QR, that the indefinite article a can also denote what
some denotes, that European and language combine via Predicate
Modification, and that speaks is a transitive verb of type ⟨e, ⟨e, t ⟩⟩.
Exercise 3. Some linguist offended every philosopher is ambiguous; it can mean either that there was one universally offensive
linguist or that for every philosopher there was a linguist, and
there may have been different linguists for different philosophers.
Explain how both of these readings can be derived using QR.
6.2.3 Generalized Argument Raising
Now we will consider a flexible types approach to the problem
of quantifiers in object position, one which repairs the type mismatch in situ (‘in place’) through type-shifting. One of the most
clever flexible types approaches is Hendriks’s (1993) system, which
can (among other things) turn a type ⟨e, ⟨e, t ⟩⟩ predicate into one
that is expecting a quantifier for its first or second argument, or
both. The instantiation of this schema that applies to quantifiers
in object position looks like this:
Variables
139
Type-Shifting Rule 2. Object raising
If α ↝ α′ where α′ is of type ⟨e, t ⟩ then:
α ↝ λv ⟨⟨e,t ⟩,t ⟩ λy [v (λz [α(z )( y )])]
as well, where Q is a variable of type ⟨⟨e, t ⟩, t ⟩ (unless Q, y or z is
free in α′ ; in that case, use different variables).
So a sentence like Sam likes everybody will be analyzed as follows:
t
(14)
∀x LOVES(x )(S)
⟨e, t ⟩
e
λy e [∀x
S
LOVES (x )( y )]
Sam
⟨⟨⟨e, t ⟩, t ⟩, ⟨e, t ⟩⟩
λv ⟨⟨e,t ⟩,t ⟩ λy e [v (λz e [LOVES(z )( y )])]
⟨⟨e, t ⟩, t ⟩
λP [∀xP (x )]
⟨e, ⟨e, t ⟩⟩
everybody
LOVES
loves
Note that this is not all that Hendriks’s system can do. He defines a general schema of which the Object Raising rule is a particular instance. The general schema is as follows: If an expression
→
→
→
has a translation α of type ⟨Ð
a , ⟨b, ⟨Ð
c , t ⟩⟩⟩, where Ð
a is a possibly
null sequence of types, then that expression also has translations
of the following form:
→
Ð
→ → [v (λz [α(Ð
→
→
→
λÐ
xÐ
x )(z )(Ð
y )])]
b
a λv ⟨⟨b,t ⟩,t ⟩ λ y Ð
c
140
Variables
A sentence with two quantifiers like Everybody likes somebody can
be analyzed in two different ways, depending on the order in which
the argument slots of the verb are lifted. The resulting type is the
same: something that combines with two quantifiers. But the different orders in which the type-shifting operations applied give
two different scopes.
It turns out that if you raise the object first,3 and then the subject,4 then you get a reading with the existential quantifier outscoping the universal quantifier. This is shown in the following tree,
where subscripts indicate the types of the variables.
(15)
t
∃ y [∀x LOVES(x )( y )]
⟨⟨e, t ⟩, t ⟩
λP [∃ yP ( y )]
⟨⟨⟨e, t ⟩, t ⟩, t ⟩
λv ⟨⟨e,t ⟩,t ⟩ [v (λz e [∀x LOVES(x )(z )])]
Somebody
⟨⟨⟨e, t ⟩, t ⟩, ⟨⟨⟨e, t ⟩, t ⟩, t ⟩⟩
λx ⟨⟨e,t ⟩,t ⟩ λv ⟨⟨e,t ⟩,t ⟩ [v (λz e [x (λz e′ [LOVES(z ′ )(z )])])]
⇑
⟨e, ⟨⟨⟨e, t ⟩, t ⟩, t ⟩⟩
λx e λv ⟨⟨e,t ⟩,t ⟩ [v (λz e [LOVES(x )(z )])]
⇑
⟨e, ⟨e, t ⟩⟩
⟨⟨e, t ⟩, t ⟩
λP [∀xP (x )]
everybody
LOVES
loves
If the subject is raised first,5 and then the object,6 then the universal quantifier ends up scoping outside of the existential quantifier.
3
→
→
with Ð
a as e, Ð
c as null, and b as e
Ð
→
→
with a as ⟨⟨e, ⟨e, t ⟩⟩, Ð
c as null, and b as e
Ð
→
Ð
→
5
with a as e, c as null, and b as e
→
→
6
with Ð
a as null, Ð
c as ⟨⟨e, t ⟩, t ⟩, and b as e
4
Variables
141
(16)
t
∀x ∃ y [LOVES(x )( y )]
⟨⟨e, t ⟩, t ⟩
λP [∃ yP ( y )]
⟨⟨⟨e, t ⟩, t ⟩, t ⟩
λy ⟨⟨e,t ⟩,t ⟩ [∀x [ y (λz e′ [LOVES(x )(z ′ )])]]
Somebody
⟨⟨⟨e, t ⟩, t ⟩, ⟨⟨⟨e, t ⟩, t ⟩, t ⟩⟩
λv ⟨⟨e,t ⟩,t ⟩ λy ⟨⟨e,t ⟩,t ⟩ [v (λz e [ y (λz e′ [LOVES(z )(z ′ )])])]
⇑
⟨⟨⟨⟨e, t ⟩, t ⟩, ⟨e, t ⟩⟩
λv ⟨⟨e,t ⟩,t ⟩ λy e [v (λz e [LOVES(z )( y )])]
⇑
⟨e, ⟨e, t ⟩⟩
⟨⟨e, t ⟩, t ⟩
λP [∀xP (x )]
everybody
LOVES
loves
6.2.4 Putative arguments for the movement
Heim & Kratzer (1998) give a number of arguments in favor of the
QR approach.
Argument #1: Scope ambiguities. The QR approach delivers two
different readings for scopally ambiguous sentences like Everybody loves somebody. Heim & Kratzer (1998) claim that this cannot
be done with a flexible types approach, but they do not consider
the approach of Hendriks (1993), which does give both readings,
as just shown in the previous section.
Argument #2: Inverse linking. A somewhat more challenging
case for a flexible types approach falls under the heading of ‘inverse linking’, also known as ‘binding out of DP’. Here is a classic
example:
142
(17)
Variables
One apple in every basket is rotten.
This does not mean: ‘One apple that is in every basket is rotten’,
and most flexible types analyses, including Hendriks 1993, deliver
only that reading. The QR analysis gives the right reading starting
from the following LF:
(18)
S
DP
every basket
1
S
VP
DP
one
is rotten
NP
D
N
apple
PP
P
t1
in
Exercise 4. Derive the translation for (18) compositionally using
the rules we have introduced so far. You may need to introduce
new lexical entries.
Exercise 5. Explain why the correct reading of (17) cannot be derived via Argument Raising.
However, Barker (2005) has a flexible types analysis of binding out
of DP (granted, one that is challenging to grasp), so this does not
constitute a knock-down argument in favor of QR.
Variables
143
Argument #3: Antecedent-contained deletion Another argument
that has been made in favor of QR is based on examples like this:
(19)
I read every novel that you did.
The VP after did seems to be deleted under identity with its antecedent. But its antecedent is the VP read every novel that you
did, so the antecedent seems to contain the VP that is deleted.
Compare this to a nomal case of ellipsis:
(20)
I read War and Peace before you did.
Here the deleted VP seems to be ‘read War and Peace’.
If the quantifier phrase in (19) undergoes QR, then the antecedent VP no longer contains the deleted VP:
(21)
S
λP
NP
every
2
N′
D
S
DP
N
novel
CP
VP
C′
which1
C
that
V
DP
read
t2
I
S
NP
you
VP
V
did
VP
V
DP
read
t1
144
Variables
The antecedent VP is now identical to the elided VP except for the
index on the trace. So QR provides a nice solution to this problem.
Jacobson (1999) argues that this phenomenon does not constitute a clear argument in favor of QR, though, because it can be
handled using a mode of composition called function composition, which we have not introduced here but has a range of interesting applications.
Argument #4: Quantifiers that bind pronouns Another argument that Heim and Kratzer give in favor of QR has to do with
reflexive pronouns. When a reflexive pronoun is anaphoric to a
proper name, the sentence can be paraphrased more or less by
replacing the pronoun with its antecedent:
(22)
a.
Mary blamed herself.
b.
Mary blamed Mary.
But this is not the case when a reflexive pronoun is anaphoric to a
quantifier:
(23)
a.
Every woman blamed herself.
b.
Every woman blamed every woman.
The problem is not unique to reflexive pronouns; it has to do with
any use of a pronoun that is connected to a quantifier including
the following:
(24)
No man noticed the snake next to him.
If we treat pronouns as variables and use QR, we can easily
account for the meanings of these sentences:
Variables
145
(25)
S
λP
DP
D
N′
every
woman
1
S
DP
t1
VP
V
DP
blamed
herself 1
Assuming that pronouns are translated as variables, it is somewhat more difficult to imagine how to derive the right reading on
a flexible types approach. Suppose we have that:
(26)
[VP [V blamed ] [DP herself1 ] ] ↝ λy [BLAMED(v 1 )( y )]
If we combine this with every woman, then we will end up with a
formula that contains a free variable:
∀x [WOMAN(x ) → BLAMED(v 1 )( y )]
and this is not the right reading.
But there are a number of ways of dealing with reflexive pronouns withour QR. For example, a reflexive pronoun can be analyzed as a function that takes as an argument a relation of type
⟨e, ⟨e, t ⟩⟩ and returns a reflexive predicate (Keenan, 1987; Szabolcsi,
1987):
(27)
herself ↝ λR ⟨e,⟨e,t ⟩⟩ λx [R (x )(x )]
Applied to for example BLAMED, this will yield the property of blaming oneself, which can be combined with every woman to give the
right reading.
146
Variables
Exercise 6. Use the treatment of herself in (27) to derive a translation for every woman blamed herself.
Argument #5: The Extraction-Scope Generalization Finally, it
has been pointed out that there seem to be some syntactic constraints on QR, and these constraints seem to mirror constraints
on movement in general. For example, the quantifier every country can take wide scope in (28a) but not in (28b); the latter has
only the implausible reading implying the existence of an “international girl”.
(28)
a.
John knows a girl from every country.
b. #John knows a girl who is from every country.
Similarly, extraction of which country is possible from a prepositional phrase modifier but not a relative clause modifier:
(29)
a.
Which country does John know a girl from?
b. *Which country does John know a girl who is from?
Example (29b) is a violation of Ross’s (1968) “Complex Noun Phrase
Constraint”, one of the so-called island constraints specifying islands for extraction (syntactic environments from which extraction is impossible). The oddness of (28b) might lead one to suspect that the constraints on scope are parallel to the constraints
on wh-extraction.
Another parallel has to do with coordination. A wide scope
reading for every man is possible in (30a), but not in (30b), where
the VP is coordinated with another VP. In (30b), we have only the
implausible reading that every man has the same mother.
(30)
a.
Some woman gave birth to every man.
b. #Some woman gave birth to every man and will eventually die.
Variables
147
Similarly, extraction of whom is possible in (31a) but not (31b),
where there is coordination.
(31)
a.
I wonder whom Mary gave birth to.
b. *I wonder whom Mary gave birth to and will eventually
die.
The badness of (31b) falls under Ross’s (1968) “Coordinate Structure Constraint”. It appears that quantifiers are also subject to
some kind of coordinate structure constraint.
Winter (2001) refers to this parallel between scope and wh- extraction as the “Extraction Scope Generalization”. This correlation
might suggest that scope readings are generated via the same kind
of movement that puts wh- phrases in their place. However, scope
does not correlate completely perfectly with extraction, and QR is
not the only possible explanation for the parallels that do exist.
Summary. While there are some phenomena that might seem
prima facie to favor a QR approach, none of them constitutes a
knock-down argument in favor of it. Some may find QR easier to
understand and work with and prefer it on those grounds; others
may find direct compositionality more appealing. QR is certainly
an important idea to have a good grasp on, as it figures in so much
writing in semantics. But whether QR has any solid empirical advantages or disadvantages seems to be an open question.
6.3 Pronouns
The composition rule we used for traces can also be used for pronouns, as we shall see presently. A theory of pronouns ough to
account for the range of uses that they have, and pronouns have
both deictic uses and anaphoric uses. A deictic use is typically
accompanied by a gesture, but may also refer to whatever is most
visually or cognitively salient. For example, if someone utters the
148
Variables
following after a certain man has left the room, then he is being
used deictically.
(32)
I am glad he is gone.
When a pronoun is anaphoric, it has a linguistic antecedent. Here
is an example:
(33)
I don’t think anybody here is interested in Smith’s work. He
should not be invited.
Ideally, it would be possible to give a unified analysis of these two
uses. This might lead one to the following hypothesis:
Hypothesis 1. All pronouns refer to whichever individual is most
salient at the moment when the pronoun is processed.
The problem with this hypothesis is that there are some pronouns
that don’t refer to any individual at all. In the following examples,
there is no particular individual that the pronoun refers to:
(34)
No1 woman blamed herself1 .
(35)
Neither1 man thought he1 was at fault.
(36)
Every1 boy loves his1 mother.
(37)
the book such1 that Mary reviewed it1
So not all pronouns are referential. Note that it is sometimes said
that No woman and herself are “coreferential” in (34) but this is
strictly speaking a misuse of the term “coreferential”, because, as
Heim and Kratzer point out, “coreference implies reference.”
The pronouns in these examples are bound variables. Let us
consider the possibility that this is the rule, and other uses of pronouns can be viewed in this way:
Variables
Hypothesis 2.
149
All pronouns are bound variables.
To make it precise, let us say that a pronoun is a bound variable if
there is a node that meets the structural description for Predicate
Abstraction dominating it and its index; otherwise it is free. (This
is a shortcut giving the effect of Heim & Kratzer’s (1998) definition,
which we will not repeat here because it involves slightly different
assumptions about the framework.)
Hypothesis 2 has a number of undesirable consequences. It
would mean that for cases like (33), we would have to QR Smith to
a position where it QRs He in the second sentence somehow. It is
not completely crazy to imagine that proper names can undergo
QR, but it doesn’t seem right that QR should allow an NP to move
across a sentence boundary.
Moreover, we don’t want to treat all pronouns as bound pronouns because there are some ambiguities that depend on a distinction between free and bound pronouns. For example, in the
movie Ghostbusters, there is a scene in which the three Ghostbusters Dr. Peter Venkman, Dr. Raymond Stanz, and Dr. Egon Spengler (played by Bill Murray, Dan Akroyd, and Harold Ramis, respectively), are in an elevator. They have just started their Ghostbusters business and received their very first call, from a fancy hotel in which a ghost has been making disturbances. They have
their proton packs on their back and they realize that they have
never been tested.
(38)
Dr Ray Stantz: You know, it just occurred to me that we really haven’t had a successful test of this equipment.
Dr. Egon Spengler: I blame myself.
Dr. Peter Venkman: So do I.
There are two readings of Peter Venkman’s quip, a sympathetic
reading and an asshole reading. On the strict reading (the sympathetic reading), Peter blames himself. On the sloppy reading (the
asshole reading), Peter blames Egon. The strict/sloppy ambiguity
150
Variables
exemplified in (38) can be explained by saying that on one reading, we have a bound pronoun, and on another reading, we have a
referential pronoun. The anaphor so picks up the the property ‘x
blames x’ on the sloppy reading, which is made available through
QR thus:
(39)
S
λP
DP
I
λ1
S
DP
VP
t1
V
DP
blame
myself 1
The strict reading can be derived from an antecedent without QR:
(40)
S
DP
I1
VP
V
DP
blame
myself 1
This suggests that pronouns are sometimes bound, and sometimes free.
Heim and Kratzer’s hypothesis: All pronouns are variables, and
bound pronouns are interpreted as bound variables, and referential pronouns are interpreted as free variables.
In the following examples, the pronoun in the sentence is free:
Variables
151
(41)
S
DP
VP
She1
(42)
V
A
is
nice
S
DP
VP
John
V
DP
hates
D
NP
his1
father
But in these examples, the pronoun is bound:
(43)
]
S
λP
DP
Every boy
λ1
S
DP
t1
VP
DP
V
loves
D
NP
his1
father
152
Variables
(44)
S
λP
DP
John
λ1
S
DP
t1
VP
DP
V
hates
D
NP
his1
father
The free/bound distinction in the syntax will be reflected in
a free/bound distinction in the logical translation. The denotation of the sentences with free pronouns will depend on an assignment.
This way of treating pronouns implies a new way of looking at
assignments. They are not only useful for defining the semantics
of quantificational expressions and such, but they also represent
the discourse context. As Heim & Kratzer (1998) put it:
Treating referring pronouns as free variables implies
a new way of looking at the role of variable assignments. Until now we have asssumed that an LF whose
truth-value varied from one assignment to the next
could ipso facto not represent a felicitous, complete
utterance. We will no longer make this assumption.
Instead, let us think of assignments as representing
the contribution of the utterance situation.
Still, it is not appropriate to say She left! if your interlocutor has
no idea who she refers to. We can encode this using the following
principle:
(45)
Appropriateness Condition
A context c is appropriate for a natural language expression
Variables
153
φ only if c determines a variable assignment g c whose domain includes every variable which has a free occurrence
in the translation of φ.
(When we say that a context “determines” an assignment, we mean
that the context contains a bunch of information, among which is
the assignment.)
Note that this idea brings us very close to dynamic semantics
(Heim, 1982, 1983b,a; Kamp & Reyle, 1993; Groenendijk & Stokhof,
1990, 1991; Muskens, 1996, among others), where the meaning of
an utterance is something that depends on and updats the current discourse context. In dynamic semantics, an indefinite noun
phrase like a man introduces a new discourse referent into the
context, expanding the domain fo the assignment function in the
context, and an anaphoric pronoun or definite description picks
up on the discourse referent.
Exercise 7. Suppose that the context provides the assignment g 1 ,
and g 1 is defined as follows:
⎡ 1 → Kim ⎤
⎢
⎥
⎢
⎥
g 1 = ⎢ 2 → Kim ⎥
⎢
⎥
⎢ 3 → Sandy ⎥
⎣
⎦
Answer the following questions:
(a) What is the translation of (41)?
(b) Is c 1 appropriate for (41)?
(c) What is the semantic value of your translation for (41) with
respect to g 1 ?
(d) Suppose we are at the beginning of a story, and the assignment of the contxt is completely empty. Is this context appropriate for (41)?
7
∣
Presupposition
7.1 The definite article
Now let us consider definite descriptions like the president. According to Russell (1905), definite descriptions are very much like
the quantifiers we have just analyzed. According to Russell, The
president smokes means ‘There is exactly one president and and
he or she smokes’:
∃x [PRESIDENT(x ) ∧ ∀ y [PRESIDENT( y ) → x = y ] ∧ SMOKES(x )
So it entails that there is a president (existence) and that there
is only one (uniqueness). So the sentence is false if there are no
presidents or multiple presidents.
Strawson (1950) finds this analysis very strange. He agrees that
definite descriptions signal existence and uniqueness of something satisfying the description, but he disagrees with Russell’s
proposal that these are entailments. As discussed in the first chapter, he says:
To say, “The king of France is wise” is, in some sense
of “imply”, to imply that there is a king of France. But
this is a very special and odd sense of “imply”. “Implies” in this sense is certainly not equivalent to “entails” (or “logically implies”).
Putting it another way:
155
156
Presupposition
When a man uses such an expression, he does not
assert, nor does what he says entail, a uniquely existential proposition. But one of the conventional functions of the definite article is to act as a signal that a
unique reference is being made – a signal, not a disguised assertion.
Strawson argues for this thesis as follows:
Now suppose someone were in fact to say to you with
a perfectly serious air: The King of France is wise. Would
you say, That’s untrue? I think it is quite certain that
you would not. But suppose that he went on to ask
you whether you thought that what he had just said
was true, or was false; whether you agreed or disagreed
with what he had just said. I think you would be inclined, with some hesitation, to say that you did not
do either; that the question of whether his statement
was true or false simply did not arise, because there
was no such person as the King of France. You might,
if he were obviously serious (had a dazed, astray-inthe-centuries look), say something like: I’m afraid you
must be under a misapprehension. France is not a monarchy. There is no King of France.
Strawson’s observation is that we feel squeamish when asked to
judge the whether a sentence of the form The F is G is true or false,
when there is no F. We do not feel that the sentence is false; we
feel that the question of its truth does not arise, as Strawson put
it. The sentence is neither true nor false. One way of interpreting
this idea is by introducing a third truth value: Along with ‘true’
and ‘false’, we will have ‘undefined’ or ‘nonsense’ as a truth value.
Let us use m as a symbol for this undefined truth value. If there is
no king of France, then the truth value of the sentence ‘The king
of France is bald’ will be m. Then the question becomes how we
can set up our semantic system so that this is the truth value that
gets assigned to such a sentence.
Presupposition
157
Intuitively, the reason that the sentence is neither true nor
false is that there is an attempt to refer to something that does
not exist. This is an intuition that was expressed earlier by Frege
(1948). According to Frege, a definite description like this, like a
proper name, denotes an individual (type e). Regarding the negative square root of 4, Frege says:
We have here a case in which out of a concept-expression,
a compound proper name is formed, with the help of
the definite article in the singular, which is at any rate
permissible when one and only one object falls under
the concept. [emphasis added]
We assume that by “concept-expression”, Frege means an expression of type ⟨e, t ⟩, and that “compound proper name”, Frege means
“a complex expression of type e”. To flesh out Frege’s analysis of
this example further, Heim and Kratzer suggest that square root is
a “transitive noun”, with a meaning of type ⟨e, ⟨e, t ⟩⟩, and that “of is
vacuous, square root applies to 4 via Functional Application, and
the result of that composes with negative under predicate modification.” This suggests the following structure:
NP: e
N′ : ⟨e, t ⟩
D: ⟨⟨e, t ⟩, e ⟩
the
A: ⟨e, t ⟩
negative
N′ : ⟨e, t ⟩
N: ⟨e, ⟨e, t ⟩⟩
square root
PP: e
P: ⟨e, e ⟩
NP: e
of
four
158
Presupposition
What does Frege mean by “permissible”? One reasonable interpretation is that the denotes a function of type ⟨⟨e, t ⟩, e ⟩ that is
only defined for input predicates that characterize one single entity. In other words, the definite article presupposes existence and
uniqueness. One might interpret this to mean that the king of
France does not denote an individual if there is no King of France,
even though it is an expression of type e. One way of handling this
is to introduce a special ‘undefined individual’ of type e. We will
use the symbol me for this.1
In order to express Frege’s idea, we need to introduce a new
symbol into the logic that we are translating English into, namly
ι ‘iota’. The idea is that ιx [P (x )] denotes the unique individual
satisfying P if there is exactly one such individual, and is otherwise
undefined. To add this symbol to our logic, first we add a syntax
rule specifying that ιx [φ] is an expression of type e:
(46)
L λ syntax rule for ι
If φ is an expression of type t , and u is a variable of type a,
then ιx [φ] is an expression of type a.
Then we add a rule specifying its semantics:
(47)
L λ semantics rule for ι
M ,g [u ↦k ]
d if {k ∶ ⟦φ⟧
= 1} = {d }
⟦ιu [π]⟧ = {
me otherwise
Now we can give a Fregean analysis of the definite determiner as
follows:
(48)
the ↝ λP [ιx [P (x )]]
Applied to a predicate-denoting expression like PRESIDENT, it denotes the unique president, if there is one and only one president
in the relevant domain.
1
Other notations that have been used for the undefined individual include
Kaplan’s (1977) †, standing for a ‘completely alien entity’ not in the set of individuals, Landman’s (2004) 0, and Oliver & Smiley’s (2013) O, pronounced ‘zilch’.
Presupposition
(49)
(50)
159
NP: e
ιx [PRESIDENT(x )]
N′ : ⟨e, t ⟩
D: ⟨⟨e, t ⟩, e ⟩
λP [ιx [P (x )]]
PRESIDENT
the
president
⟦ιx [PRESIDENT(x )]⟧M ,g
M ,g [x ↦k ]
d if {k ∶ ⟦PRESIDENT⟧
= 1} = {d }
={
me otherwise
For example, in a model corresponding to the White House in
2014, where the only president in the relevant domain is Barack
Obama, the denotation of the phrase would be Barack Obama.
Now, let us assume that a predicate like BALD only yields true
or false for actual individuals, and yields the undefined truth value
m when given the undefined individual me as input. Since the
King of France (used today) would have me as its denotation, the
denotation of The King of France is bald would then have m as its
denotation.
Exercise 8. Suppose that your friend doesn’t understand why this
is true. Explain it to him or her.
Let us consider another example. Beethoven wrote one opera,
namely Fidelio, but Mozart wrote quite a number of operas. So
in a model reflecting this fact of reality, the phrase the opera by
Beethoven has a defined value. But the opera by Mozart does not.
Consider what happens when the opera by Mozart is embedded in
a sentence like the following:
(51)
The opera by Mozart is boring.
160
Presupposition
This might have the following translation:
BORING (ιx [ OPERA (x ) ∧ BY ( MOZART )(x )])
But this expression will not have a value in a model where there
are multiple operas by Mozart. In this way, the undefinedness of
the definite description “percolates up”, as it were, to the sentence
level.
Exercise 9. Give a tree showing how the truth and definedness
conditions of The opera by Mozart is boring are built up compositionally following the assumptions given in this section.
7.2 Possessives
It is often suggested that possessive descriptions contain a hidden
the. One prima facie reason for believing this is that possessive descriptions with definite possessors in argument position typically
behave like (constant) terms, i.e., behave as if they have denotations of type e. For example, Mary’s rabbit gives rise to a contradiction in (52) whereas some rabbit does not. This would be expected
if Mary’s rabbit denoted a single rabbit.
(52)
a.
Mary’s rabbit is in the cage and Mary’s rabbit is outside
the cage. (contradictory)
b.
Some rabbit is in the cage and some rabbit is outside
the cage. (not contradictory)
Mary’s rabbit also behaves like an individual-denoting expression
when it comes to conjunction.
(53)
a.
Mary’s rabbit is from South Africa and Mary’s rabbit is
a secret agent.
⇐⇒ Mary’s rabbit is a secret agent from South Africa.
Presupposition
b.
161
Some rabbit is from South Africa and some rabbit is a
secret agent.
⇐⇒
/ Some rabbit is a secret agent from South Africa.
However, as Löbner (2011) discusses, these tests do not always indicate type e status when the possessor is not definite.
(54)
At least one child’s pet rabbit is from South Africa and at
least one child’s pet rabbit is a secret agent.
⇐⇒
/ At least one child’s pet rabbit is a secret agent from
South Africa.
(55)
At least one child’s pet rabbit isn’t in the cage, and at least
one child’s rabbit is in the cage. (not contradictory)
Does this show that possessive phrases with quantificational determiners are not of type e? Perhaps not. As Kamp (to appear)
observes, at least one child’s pet rabbit is felicitous only when at
least one child in the domain of discourse has exactly one pet rabbit. This suggests that possessives can, for the purposes of these
kinds of examples, be treated as definite within the scope of the
quantifiers they contain.
In order to analyze these cases, we need to have some kind
of analysis of possessives. It is a bit tricky, because in a case like
John’s sister, the possessive relation seems to come from the noun,
whereas in John’s car, the possessive relation comes from context;
John could own the car, or have been assigned to the car, etc.
Nouns like sister are called relational nouns, and are typically analyzed as being of type ⟨e, ⟨e, t ⟩⟩. So sister would be translated as
a two-place predicate. One way to distinguish between relational
and non-relational nouns is with of-PP complements as in (56a)
vs. the so-called double genitive construction shown in (56b).
(56)
a.
a portrait of Picasso
b.
a portrait of Picasso’s
The example without possessive ’s in (56a) describes a portrait in
which Picasso is depicted. In (56b), Picasso could have made the
162
Presupposition
portrait, or he could own it, or any number of other things. The
restricted interpretation of (56a) can be explained based on the
following assumptions. There is one sense of portrait on which it
denotes a binary predicate, and on that sense, the argument to the
noun is the thing depicted in the portrait. On that sense, portrait
is a transitive noun, and requires an object, just as transitive verbs
require an object, and the of-PP construction in (56a) serves to
feed the object to the noun. So the of-PP construction in (56a)
is not a possessive construction at all; it is simply a construction
in which a transitive noun takes an argument. There is another
sense of portrait which is a simple unary predicate, and that is
the sense that surfaces in (56b), where the possessive ’s invokes a
contextually salient possessive relation.
To deal with this, we will take inspiration from Vikner & Jensen’s
(2002) idea that in cases like John’s car, an ⟨e, t ⟩-type noun undergoes a type shift to become an ⟨e, ⟨e, t ⟩⟩-type noun whose meaning introduces an underspecified possessive relation POSS. The
type-shift is defined as follows:
Type-Shifting Rule 3. Possessive shift
If α ↝ α′ where α is of type ⟨e, t ⟩, then:
α ↝ λyλx [α′ (x ) ∧ POSS( y )(x )]
as well (unless y or x is free in α′ ; in that case, use different variables of the same type.
For example, this type-shift applied to CAR will produce
λyλx [CAR(x ) ∧ POSS( y )(x )]
so the possessive shift turns car into car of, so to speak.
One of the virtues of the possessive shift is that it makes it possible to account for the ambiguity of examples involving former
such as the following:
Presupposition
(57)
163
John’s former mansion
‘the mansion that John formerly owned’
‘the former mansion that John currently owns’
If the possessive shift occurs between former and mansion, then
we get one reading, but if it occurs above former mansion, then
we get another reading. A full analysis of this particular example
would require us to be able to talk about time, which we have so
far ignored, so let us move on without getting into detail about
that.
We will also assume that possessive ’s is an identity function
on binary predicates, so that any noun that it combines with has
to undergo the possessive shift.2
(58)
’s ↝ λR [R ]
(Recall that R is a variable of type ⟨e, ⟨e, t ⟩⟩.) The possessive ’s will
form a constituent with the modified noun. This may strike the
reader as odd, but it is not so strange in light of other constructions involving clitics. (Clitics are things that are in-between a
morpheme and a word.) For example, in I’m tired, it is not unreasonable to say that ’m forms a syntactic constient with tired.
(Barker (1995) assumes that the possessive construction involves
a silent head that combines with the following noun and serves
as “glue” for the possessive construction; this analysis assumes
that the possessive ’s sits in the position where Barker has a silent
head.) So the types for John’s rabbit will work out as follows:
2
Vikner & Jensen (2002) following Partee (1983/1997), assume a more complicated analysis of ’s.
164
Presupposition
DP: ⟨e, t ⟩
D′ : ⟨e, ⟨e, t ⟩⟩
DP: e
John
D: ⟨⟨e, ⟨e, t ⟩⟩, ⟨e, ⟨e, t ⟩⟩⟩
NP: ⟨e, ⟨e, t ⟩⟩
’s
⇑
⟨e, t ⟩
rabbit
Notice that we have used DP (“Determiner Phrase”) instead of NP
in the syntax here. This is because it is convenient to use the specifier position in the DP (the sister of D′ ) for the possessor phrase
(Abney, 1987). For consistency, we ask the reader to mentally convert all instances of:
NP
D
N′
determiner
N
noun
in the foregoing text to:
DP
D′
D
NP
determiner
N′
N
noun
Presupposition
165
or to consider the former an abbreviation for the latter.
To get a type e interpretation of John’s rabbit, we can assume
that there is a silent definite article in the syntax, or we can assume
that John’s rabbit undergoes a type-shifting operation that does
the same job. In fact, Partee (1986) already posited a type-shifting
operation that would do this job, called IOTA.
Type-Shifting Rule 4. The IOTA shift
If α ↝ α′ where α′ is of type ⟨e, t ⟩, then
α ↝ ιx [α′ (x )]
as well (unless x is free in α′ ; then choose a different variable).
If we assume that this type-shifting operation applies to John’s rabbit, then we will end up with the following translation for the phrase:
(59)
ιx [POSS(x )(J) ∧ RABBIT(x )]
Exercise 10. Give a tree analysis of John’s rabbit showing logical
representations at each node. At each node, indicate which composition rules and type-shifting operations you use, if any.
Now let us consider how to analyze a case with a quantified
possessor. We can do it with Quantifier Raising. The possessor,
which is a quantifier, undergoes QR, leaving a trace in its original
position.3 The possessive noun undergoes the IOTA shift, yielding,
in effect, a definite description containing a variable bound by the
possessor. The syntactic structure and types will be as follows:
3
This is an instance of the ‘binding out of DP’ problem, so QR is making our
lives easier here. See Barker (2005) and Francez (2011) regarding directly compositional analyses of quantified possessives.
166
Presupposition
(60)
S
t
λP
DP
⟨⟨e, t ⟩, t ⟩
some child
1
S
t
VP
DP
e
⟨e, t ⟩
⇑
⟨e, t ⟩
is a secret agent
D′
⟨e, ⟨e, t ⟩⟩
DP
e
t1
D
NP
⟨⟨e, ⟨e, t ⟩⟩, ⟨e, ⟨e, t ⟩⟩⟩
⟨e, ⟨e, t ⟩⟩
⇑
⟨e, t ⟩
’s
pet rabbit
Exercise 11. Derive the translation for Some child’s pet rabbit is a
secret agent step-by-step.
7.3 Definedness conditions
The definite article is one of many presupposition triggers, and
even if we are satisfied with our treatment of the definite article,
we still need a more general way of dealing with presupposition.
Presupposition
167
The determiners both and neither, for example, come with presuppositions. In a context with three candidates for a job, it would
be quite odd for someone to say either of the following:
(61)
a.
Both candidates are qualified.
b.
Neither candidate is qualified.
If there are two candidates and both are qualified, then (61a) is
clearly true and (61b) is clearly false. But if there is any number
of candiates other than two, then it is hard to say whether these
sentences are true or false. This suggests that both candidates and
neither candidate come with a presupposition that there are exactly two candidates.
As the reader may have noticed, we have been assiduously
avoiding plural noun phrases so far, so we will ignore both and focus on neither. We can model this presupposition by treating neither as a variant of every that is only defined when its argument is a
predicate with exactly two satisfiers. Let us use ∣P ∣ = 2 (suggesting
‘the cardinality of P is 2’) as a way of writing the idea that predicate P has exactly two satisfiers.4 This is what is presupposed.
To signify that it is presupposed, we will use Beaver & Krahmer’s
(2001) ∂ ‘partial’ operator. A formula like this:
∂[∣P ∣ = 2]
can be read, ‘presupposing that there are exactly two P s’. The lexical entry for neither can be stated using the ∂ operator as follows:
(62)
neither ↝ λP λQ [∂[∣P ∣ = 2] ∧ ¬∃x [P (x ) ∧ Q (x )]]
This say that neither is basically a synonym of no, carrying an extra
presupposition: that there are exactly two P s.
In order to be able to give translations like this, we need to
augment L λ to handle formulas containing the ∂ symbol. So ∂[φ]
will be a new kind of expression of type t in our logic. Its value will
4
∣P ∣ = 2 is short for ∃x ∃ y [x ≠ y ∧ P (x ) ∧ P ( y ) ∧ ¬∃z [z ≠ x ∧ z ≠ y ∧ P (z )]].
168
Presupposition
be ‘true’ if φ is true and ‘undefined’ otherwise. To implement this,
we must add a new clause to our syntax rules for L λ :
(63)
Syntax of definedness conditions
If φ is an expression of type t , then ∂[φ] is an expression of
type t .
We define the semantics of these expressions as follows:
(64)
Semantics of definedness conditions
If φ is an expression of type t , then:
⟦∂[φ]⟧M ,g = {
⟦α⟧M ,g if ⟦φ⟧M ,g = 1
m otherwise.
The lexical entry in (62) will give us the following analysis for
(61b), where β-reduced variants of the translations are given at
each node:
(65)
∂[∣CANDIDATE∣ = 2] ∧ ∀x [CANDIDATE(x ) → QUALIFIED(x )]
λQ [∂[∣CANDIDATE∣ = 2] ∧ ∀x [CANDIDATE(x ) → Q (x )]]
QUALIFIED
is qualified
λP λQ [∂[∣P ∣ = 2] ∧ ∀x [P (x ) → Q (x )]]
CANDIDATES
neither
candidate
The translation for the whole sentence (at the top of the tree) should
have a defined value in a model if ∣CANDIDATE∣ = 2 is true in the
model. If it has a defined value, then its value is equal to that of
∀x [CANDIDATE(x ) → QUALIFIED(x )].5
5
Notice that β-reduction works as usual under this way of doing things. Although the notation here is quite similar to Heim & Kratzer’s (1998) notation for
partial functions, and the expressive power is the same, the style of handling undefinedness laid out in Heim & Kratzer (1998) does not allow for β-reduction to
work properly, so the present system is a bit cleaner in this respect.
Presupposition
169
The quantifier every is also sometimes argued to come with
a presupposition. For example, one might hesitate to judge the
following sentence as true:
(66)
Every unicorn in the world is white.
Assuming that there are no unicorns in the world, this sentence
is true given the analysis of every that we have given so far: since
there are no unicorns in the world, there are no unicorns in the
world that are not white, and this is sufficient to make the sentence true according to the semantics we have given. But something feels wrong with (66), suggesting that there might be something wrong with our analysis of every. It seems to imply, in some
sense of imply (as Strawson would say), that there are unicorns.
But (66) does not entail that there are unicorns; we would hesitate
to judge it as true or false. We can capture this using the following
kind of analysis of every:
(67)
every ↝ λP λQ [∂[∃xP (x )] ∧ ∀x [P (x ) → Q (x )]]
This will give rise to an undefined value for (66) in models where
there are no unicorns (such as the one corresponding to reality),
capturing the intuition that the sentence is neither true nor false.
7.4 Projection problem
This treatment of presupposition captures the fact, discussed in
the first chapter, that presuppositions project. If there is no king
of France, then The king of France is wise has no truth value, nor
does The king of France is not wise. This captures the intuition that
both The king of France is wise and The king of France is not wise
imply – in some sense of imply – that there is a king of France.
But presuppositions do not always project, as Karttunen (1973)
discussed. Consider the following examples:
(68)
If there is a king of France, then the king of France is wise.
170
Presupposition
(69)
Either there is no king of France or the king of France is
wise.
Neither of these sentences as a whole implies that there is a king
of France. In Karttunen’s terms, if/then and either/or are filters,
which do not let all presuppositions “through”, so to speak. Imagine the presuppositions floating up from deep inside the sentence,
and getting trapped when they meet if/then or either/or. The problem of determining when a presupposition projects is called the
projection problem.
Operators like if/then and either/or do let some presuppositions through, for example:
(70)
If France is neutral, then the king of France is wise.
(71)
Either France is lucky or the king of France is wise.
Karttunen gave the following generalization: When the antecedent
of the conditional (the if-part) entails a presupposition of the consequent, the presupposition gets filtered out. In (68), the consequent (the king of France is wise) presupposes that there is a king
of France, and the antecedent of the conditional is there is a king
of France. The antecedent entails of course that there is a king
of France, so the presupposition gets filtered out. In (70), the antecedent is France is staying out of the war, which doesn’t entail
that there is a king of France, so the presupposition “passes through
the filter”, so to speak.
With a disjunction, the generalization is as follows: a presupposition of one disjunct gets filtered out when the negation of another disjunct entails it, as in (69). In (69), the second disjunct (the
king of France is wise) presupposes that there is a king of France.
The first disjunct is there is no king of France, whose negation is
there is a king of France, which again entails of course that there
is a king of France, so the presupposition gets filtered out. In (71),
the first disjunct does not entail that there is a king of France, so
the presupposition does not get filtered out.
Presupposition
171
These two generalizations can be stated concisely and illuminatingly using Karttunen’s (1974) concept of local context: In general, a presupposition gets filtered out if it is entailed by the appropriate local context. The local context for the consequent of
a conditional is its antecedent, and the local context for one disjunct of a disjunction is the negation of the other disjunct.
Note that in (68) and (69), the local context is the same as the
presupposition (namely that there is a king of France), but it need
not be that way. Consider the following example from Karttunen
(1973):
(72)
Either Geraldine is not a mormon or she has stopped wearing her holy underwear.
The second disjunct (she has stopped wearing her holy underwear)
presupposes that Geraldine has holy underwear. The local context for the second disjunct is the negation of the first disjunct.
The first disjunct is Geraldine is not a mormon, so the local context is Geraldine is a mormon. If we assume that all mormons have
holy underwear, then the local context entails that Geraldine has
holy underwear. So Karttunen’s generalization correctly captures
the fact that (72) does not presuppose that Geraldine has holy underwear.
The system that we have introduced for dealing with presuppositions might seem to predict that presuppositions will always
project, since undefinedness tends to “percolates up,” so to speak.
The projection problem has been dealt with elegantly using dynamic semantics, where the meaning of a sentence is a “context
change potential”: a function that can update a discourse context. See for example Heim (1983b) and Beaver (2001) to learn
more about this. However, Beaver & Krahmer (2001) argue that
presupposition projection can in fact be handled well in a static
semantics with three truth values (true, false and undefined), using the ∂ operator, as we have done here.
8
∣
Summary
We are almost done with all of English, but not quite. Ha! There
are extremely many topics which are fruitful to study from this
perspective that we haven’t touched on at all:
• tense: John loved Mary vs. John loves Mary
• aspect: John is kissing/*loving Mary
• modality: John might/must/should kiss Mary
• attitude reports: John believes that he is kissing Mary
• conjoined and disjoined noun phrases: John and Mary, John
or Mary
• plurals: book vs. books
• numerals: two books, my two books, the two books
• comparatives: prettier, more beautiful, more books, less pretty,
fewer books, less milk
• superlatives: prettiest, most pretty, most books
• exclusives: only, sole(ly), exclusive(ly), mere(ly), just
• exceptives: except (for), save (that), but
• indexicals: I, you, here, now, today
173
174
Summary
• demonstratives: that glass over there
• questions: Did John kiss Mary? and embedded questions:
John doesn’t know whether he kissed Mary
• imperatives: Kiss Mary!
to name a few. And there is much remaining to be said about the
topics we have touched on. However, the reader now has a starter
kit. The following sections give the fragment of English that we
have developed so far.
8.1 Logic: Partial typed lambda calculus (L3 )
Expressions of the following fragment of English given below will
be translated into the following version of lambda calculus in which
there are three truth values. Let us call the language L3 .
Types. e and t are types, and if a and b are types, then ⟨a, b ⟩ is
a type; nothing else is a type. For all type a, ma stands for the
undefined entity of type a.
8.1.1 Syntax of L3
The set of expressions of type a, for any type a, is defined recursively as follows. (An expression of type t is a formula.)
1. Basic expressions
For each type a,
(a) the non-logical constants of type a is the set of symbols of the form c n,a for each natural number n.
(b) the variables of type a is the set of symbols of the form
v n,a for each natural number n.
Summary
175
2. Predication
If α and β are expresions of type ⟨a, b ⟩ and a respectively,
then α(β) is an expression of type b.
3. Negation
If φ is a formula, then so is ¬φ.
4. Binary Connectives
If φ and ψ are formulas, then so are [φ ∧ ψ], [φ ∨ ψ], [φ → ψ]
and [φ ↔ ψ].
5. Quantification
If φ is a formula and u is a variable (of any type), then ∀uφ
and ∃uφ are formulas.
6. Lambda abstraction
If α is an expression of type a adn u is a variable of type b,
then λu [α] is an expression of type ⟨b, a ⟩.
7. Iota terms
If φ is a formula, and u is a variable of type a, then ιu [φ] is
an expression of type a.
8. Definedness conditions
If φ is a formula, then ∂[φ] is a formula.
8.1.2 Semantics of L3
For each type a, there is an associated domain D a . D e is the domain of individuals, D t is the set of truth values, and for any types
a and b, D ⟨a,b ⟩ is the set of functions from D a to D b .
Expressions are interpreted in L3 with respect to both:
• a model M = ⟨D, I ⟩ where D is a non-empty set of individuals, and I is a function assigning a denotation in D a to each
non-logical constant of type a
176
Summary
• an assignment g , which is a function assigning to each variable of type a a denotation from the set D a
For every well-formed expression α, the semantic value of α with
M ,g
respect to model M and assignment function g , written ⟦α⟧
, is
defined recursively as follows:
1. Basic expressions
(a) If α is a non-logical constant, then ⟦α⟧
(b) If α is a variable, then ⟦α⟧
M ,g
M ,g
= I (α).
= g (α).
2. Predication
For any types a and b, if α is an expression of type ⟨a, b ⟩ and
β is an expression of type a, then ⟦α(β)⟧ = ⟦α⟧(⟦β⟧).
3. Negation
M ,g
If φ is a formula, then ⟦¬φ⟧
= 1 iff ⟦φ⟧M ,g = 0.
4. Binary connectives
If φ and ψ are formulas, then:
(a) ⟦φ ∧ ψ⟧
M ,g
= 1 iff ⟦φ⟧M ,g = 1 and ⟦ψ⟧M ,g = 1.
(b) ⟦φ ∨ ψ⟧
M ,g
= 1 iff ⟦φ⟧M ,g = 1 or ⟦ψ⟧M ,g = 1.
(c) ⟦φ → ψ⟧
M ,g
= 1 iff ⟦φ⟧M ,g = 0 or ⟦ψ⟧M ,g = 1.
M ,g
= 1 iff ⟦φ⟧M ,g = ⟦ψ⟧M ,g .
(d) ⟦φ ↔ ψ⟧
5. Quantification
If φ is a formula and u is a variable of type a then:
(a) ⟦∀uφ⟧
M ,g
= 1 iff for all k ∈ D a :
⟦φ⟧M ,g [u ↦k ] = 1.
(b) ⟦∃uφ⟧
M ,g
= 1 iff there is a k ∈ D a such that:
⟦φ⟧M ,g [u ↦k ] = 1.
Summary
177
6. Lambda abstraction
If α is an expression fo type a and u is a variable of type b
M ,g
then ⟦λu [α]⟧
is that function h from D b into D a such
M ,g [u ↦k ]
that for all objects k in D b , h (k ) = ⟦α⟧
.
7. Iota terms
If φ is a formula and u is a variable of type a then:
⟦ιu [φ]⟧ = {
M ,g [u ↦k ]
d if {k ∶ ⟦φ⟧
#e otherwise.
= 1} = {d }
8. Definedness conditions
If φ is an expression of type t , then:
⟦∂[φ]⟧M ,g = {
⟦α⟧M ,g if {⟦φ⟧M ,g = 1}
ma otherwise.
M
M ,g
Truth in a model. For any expression φ, ⟦φ⟧ = 1 iff ⟦φ⟧
=1
M
M ,g
for every value assignment g . Similarly, ⟦φ⟧ = 0 iff ⟦φ⟧
= 0 for
every value assignment g .
8.2 Syntax of English fragment
Syntax rules.
The following rules derive trees at Deep Structure:
S
S
JP
VP
AP
DP
D′
NP
N′
N′
→
→
→
→
→
→
→
→
→
→
DP VP
S JP
JS
V (DP∣AP∣PP∣CP)
A (PP)
(DP) D′
D (NP)
D (N′ )
N (PP∣CP)
A N′
178
Summary
PP
CP
C′
→
→
→
P DP
C′
CS
Lexicon. Lexical items are associated with syntactic categories
as follows:
and, or
it is not the case that
smokes, loves, kissed, is
lazy, proud
drunkard, baby, kid, zebra, sister
the, a, every, some, no, neither, ’s,
J:
Neg:
V:
A:
N:
D:
who, which
John, Obama, everybody, somebody, nobody...
of, with
that
P:
C:
Transformations. We assume the ‘T-model’, where a set of transformations convert Deep Structures to Surface Structures, Surface
Structures to Phonological Forms, and Surface Structures to Logical Forms.
DS
SS
LF
PF
The only transformation from Deep Structure to Surface Structure
that we will make explicit here is Relativization (cf. Muskens 1996):
Relativization.
If α is a DP dominating who, whom or which:
[CP [C′ X α ]Y ] ] ⇒ [CP αi [C′ X [ t i Y ] ] ]
Summary
179
where i is a fresh index.
The structures that are interpreted are Logical Forms, which
are derived from Surface Structures using Quantifier Raising (QR).
Following May (1985), we assume that QR only allows adjunction
to S nodes (whereas Heim & Kratzer (1998) allow adjunction to
any expression of an appropriate semantic type), but we take the
insertion of a numerical index into the tree from Heim & Kratzer
(1998).
Quantifier Raising.
If α is a DP:
[S X α Y ] ⇒ [ α [ i [S X [DP t i ] Y ] ] ]
where i is a fresh index.
8.3 Translations
8.3.1 Lexical entries
We associate each lexical item with a translation to L3 . We will use
the following abbreviations:
• x is v 0,e , y is v 1,e , and z is v 2,e .
• X , Y , P and Q are variables of type ⟨e, t ⟩.
• R is a variable of type ⟨e, ⟨e, t ⟩⟩.
• p and q are variables of type t .
and R are constants of type e.
•
B, L, M, H
•
DRUNKARD, BABY KID, ZEBRA , LAZY , and SNORES are constants
of type ⟨e, t ⟩.
•
LOVES , KISSED, WITH , PROUD , and SISTER are constants of type
⟨e, ⟨e, t ⟩⟩.
180
Summary
Type ⟨e, t ⟩:
1. drunkard ↝ DRUNKARD
2. baby ↝ BABY
3. kid ↝ KID
4. zebra ↝ ZEBRA
5. lazy ↝ LAZY
Type e:
1. Homer ↝ H
2. Maggie ↝ G
3. Bart ↝ B
4. Lisa ↝ L
5. Marge ↝ M
Type ⟨t , ⟨t , t ⟩⟩:
1. and ↝ λpλq [p ∧ q ]
2. or ↝ λpλq [p ∨ q ]
Type ⟨t , t ⟩:
1. it is not the case that ↝ λp [¬p ]
Type ⟨⟨e, t ⟩, ⟨e, t ⟩⟩:
1. is↝ λP [P ]
2. a ↝ λP [P ]
Type ⟨⟨e, t ⟩, e ⟩:
Summary
1. the ↝ λP [ιx [P (x )]]
Type ⟨e, ⟨e, t ⟩⟩:
1. loves ↝ LOVES
2. kissed ↝ KISSED
3. with ↝ WITH
4. proud ↝ PROUD
5. sister ↝ SISTER
Type ⟨⟨e, t ⟩, t ⟩:
1. something ↝ λP [∃x [P (x )]]
2. nothing ↝ λP [¬∃x [P (x )]]
3. everything ↝ λP [∀x [P (x )]]
Type ⟨⟨e, t ⟩, ⟨⟨e, t ⟩, t ⟩⟩:
1. some ↝ λP λQ [∃x [P (x ) ∧ Q (x )]]
2. no ↝ λP λQ [¬∃x [P (x ) ∧ Q (x )]]
3. every ↝ λP λQ [∂[∃x [P (x )]] ∧ ∀x [P (x ) → Q (x )]]
4. neither ↝ λP λQ [∂[∣P ∣ = 2] ∧ ¬∃x [P (x ) ∧ Q (x )]]
Type ⟨e, e ⟩:
1. of ↝ λx [x ]
181
182
Summary
8.3.2 Composition rules
If the translation of an expression γ is not specified in the lexicon,
then it is given by one of the following rules:
1. Functional Application
If α ↝ α′ , where α′ is type ⟨a, b ⟩ and β ↝ β′ , where β′ is
of type a, and γ is an expression whose only two subconstituents are α and β, then:
γ ↝ α′ (β′ )
2. Predicate Modification
If α ↝ α′ and β ↝ β′ and α′ and β′ are of type ⟨e, t ⟩, and γ is
an expression whose only two subconstituents are α and β,
then:
γ ↝ λu [α′ (u ) ∧ β′ (u )]
where u is a variable of type e that does not occur free in α′
or β′ .
3. Predicate Abstraction
If β ↝ β′ where β′ is an expression of type t , and γ is an
expression whose only two subconstituents are αi and β,
then:
γ ↝ λv i ,e [β′ ]
4. Pronouns and Traces
If α is a trace or a pronoun, then:
αi ↝ v i ,e
5. Non-branching Nodes
If α ↝ α′ , and β is an expression whose only daughter is α,
then
β ↝ α′
Summary
183
We also have the following type-shifting rules:
1. Predicate-to-modifier shift
If α ↝ α′ , where α′ is of category ⟨e, t ⟩, and π is a variable of
type ⟨e, t ⟩ that is not free in α′ , then
α ↝ λπλx [α′ (x ) ∧ π(x )]
as well (as long as P and x are not free in α′ ; in that case,
use different variables of the same type).
2. Argument Raising
→
→
If an expression has a translation α of type ⟨Ð
a , ⟨b, ⟨Ð
c , t ⟩⟩⟩,
then that expression also has translations of the following
form:
→
→
Ð
→ → [v (λz [α(Ð
→
→
x )(z )(Ð
y )])]
λÐ
xÐ
b
c
a λv ⟨⟨b,t ⟩,t ⟩ λ y Ð
3. Possessive shift
If α ↝ α′ where α is of type ⟨e, t ⟩, then:
α ↝ λyλx [α′ (x ) ∧ POSS( y )(x )]
4. Iota shift
If α ↝ α′ where α′ is of type ⟨e, t ⟩, then
α ↝ ιx [α′ (x )]
as well (unless x is free in α′ ; then choose a different variable).
Bibliography
Abney, Steven Paul. 1987. The English noun phrase in its sentential
aspect: MIT dissertation.
Barker, Chris. 1995. Possessive descriptions. Stanford: CSLI Publications.
Barker, Chris. 2005. Remark on Jacobson 1999: Crossover as a local constraint. Linguistics and Philosophy 28(4). 447–472.
Barwise, Jon & Robin Cooper. 1981. Generalized quantifiers and
natural language. Linguistics and Philosophy 4. 159–219.
Beaver, David. 2001. Presupposition and assertion in dynamic semantics. Stanford: CSLI Publications.
Beaver, David & Emiel Krahmer. 2001. A partial account of presupposition projection. Journal of Logic, Language and Information 10. 147–182.
Bittner, Maria. 1994. Cross-linguistic semantics. Linguistics and
Philosophy 17. 53–108.
Champollion, Lucas, Josh Tauberer & Maribel Romero. 2007. The
penn lambda calculator: Pedagogical software for natural language semantics. In Tracy Holloway King & Emily M. Bender
(eds.), Proceedings of the grammar engineering across frameworks (geaf ) 2007 workshop, CSLI On-line Publications.
185
186
BIBLIOGRAPHY
Chierchia, Gennaro & Sally McConnell-Ginet. 1990. Meaning and
grammar: An introduction to semantics. Cambridge, MA: MIT
Press 2nd edn.
Chomsky, Noam. 1957. Syntactic structures. The Hague: Mouton.
Chomsky, Noam. 1965. Aspects of the theory of syntax. Cambridge,
MA: MIT Press.
Chomsky, Noam & Howard Lasnik. 1977. Filters and control. Linguistic Inquiry 8. 435–504.
Condoravdi, Cleo. 2010. NPI licensing in temporal clauses. Natural Language and Linguistic Theory 28. 877–910.
Cooper, Robin. 1983. Quantification and syntactic theory. Reidel.
Dowty, David, Robert E. Wall & Stanley Peters. 1981. Introduction
to Montague semantics. Dordrecht: Kluwer.
von Fintel, Kai. 1999. NPI licensing, Strawson entailment, and
context-dependency. Journal of Semantics 16. 97–148.
Francez, Itamar. 2011. Quantified possessives and direct compositionality. In Ed Cormany, Satoshi Ito & David Lutz (eds.),
Proceedings of Semantics and Linguistic Theory (SALT), vol. 19,
165–179. eLanguage.
Frege, Gottlob. 1948. Sense and reference. The Philosophical Review 57(3). 209–230.
Groenendijk, Jeroen & Martin Stokhof. 1990. Dynamic montague
grammar. In Proceedings of the second symposium on logic and
language, 3–48. Budapest.
Groenendijk, Jeroen & Martin Stokhof. 1991. Dynamic predicate
logic. Linguistics and Philosophy 14. 39–100.
BIBLIOGRAPHY
187
Heim, Irene. 1982. The semantics of definite and indefinite noun
phrases. (Electronically searchable version available on Semantics Archive): MIT dissertation.
Heim, Irene. 1983a. File change semantics and the familiarity theory of definiteness. In Rainer Bäuerle, Christoph Schwarze &
Arnim von Stechow (eds.), Meaning, use, and the interpretation
of language, 164–189. Berlin: Walter de Gruyter.
Heim, Irene. 1983b. On the projection problem for presuppositions. In Daniel Flickinger, Michael Barlow & Michael Westcoat
(eds.), Proceedings of the second west coast conference on formal
linguistics, 114–125. Stanford, CA: Stanford University Press.
Heim, Irene & Angelika Kratzer. 1998. Semantics in generative
grammar. Oxford: Blackwell.
Hendriks, Hermann. 1993. Studied flexibility: ILLC dissertation.
Jacobson, Pauline. 1999. Towards a variable-free semantics. Linguistics and Philosophy 22. 117–184.
Jacobson, Pauline. 2000. Paycheck pronouns, Bach-Peters sentences, and variable-free semantics. Natural Language Semantics 8. 77–155.
Jacobson, Pauline. 2012. Direct compositionality. In The Oxford handbook of compositionality, 109–129. Oxford University
Press.
Kamp, Hans. to appear. The imporance of presupposition. In
Antje Rossdeutscher Christian Rohrer & Hans Kamp (eds.), Linguistic form and its computation, Stanford: CSLI Publications.
Kamp, Hans & Uwe Reyle. 1993. From discourse to logic. Dordrecht: Kluwer Academic Publishers.
188
BIBLIOGRAPHY
Kaplan, David. 1977. Demonstratives: An essay on the semantics, logic, metaphysics, and epistemology of demonstratives
and other indexicals. In John Perry Joseph Almog & Howard
Wettstein (eds.), Themes from Kaplan, 267–298. Oxford: Oxford
University Press.
Karttunen, Lauri. 1973. Presuppositions of compound sentences.
Linguistic Inquiry 4(2). 169–193.
Karttunen, Lauri. 1974. Presuppositions and linguistic context.
Theoretical Linguistics 1. 181–194.
Keenan, Edward. 1987. Semantic case theory. In Proceedings of
the sixth Amsterdam Colloquium, University of Amsterdam.
Klein, Ewan & Ivan A. Sag. 1985. Type-driven translation. Linguistics and Philosophy 8. 163–201.
Ladusaw, William A. 1980. On the notion ‘affective’ in the analysis
of negative polarity items. Journal of Linguistic Research 1. 1–
16.
Landman, Fred. 2004. Indefinites and the type of sets. Malden, MA:
Blackwell.
Löbner, Sebastian. 2011. Concept types and determination. Journal of Semantics 28. 279–333.
May, Robert. 1985. Logical form: Its structure and derivation. MIT
Press.
Montague, Richard. 1974a. English as a formal language. In Richmond H. Thomason (ed.), Formal philosophy, 188–221. New
Haven: Yale University Press.
Montague, Richard. 1974b. The proper treatment of quantification in ordinary English. In Richmond H. Thomason (ed.), Formal philosophy, New Haven: Yale University Press.
BIBLIOGRAPHY
189
Muskens, Reinhard. 1996. Combining Montague semantics and
discourse representation. Linguistics and Philosophy 19. 143–
186.
Oliver, Alex & Timothy Smiley. 2013. Plural logic. Oxford University Press.
Partee, Barbara. 1983/1997. Uniformity vs. versatility: The genitive, a case study, appendix to T. Janssen (1997) “Compositionality”. In Johan van Benthem & Alice ter Meulen (eds.), The
handbook of logic and language, Amsterdam: Elsevier.
Partee, Barbara H. 1986. Noun phrase interpretation and typeshifting principles. In Jeroen Groenendijk, Dick de Jongh &
Martin Stokhof (eds.), Studies in Discourse Representation Theory and the theory of generalized quantifiers, 115–143. Dordrecht: Foris.
Partee, Barbara H. 1995. Lexical semantics and compositionality.
In Lila Gleitman & Mark Lieberman (eds.), Invitation to cognitive science, vol. I, 211–160. Cambridge, MA: MIT Press.
Partee, Barbara H., Alice ter Meulen & Robert E. Wall. 1990. Mathematical methods in linguistics. Dordrecht: Kluwer.
Pollard, Carl & Ivan A. Sag. 1994. Head-driven phrase structure
grammar. Chicago: University of Chicago Press.
Ross, John Robert. 1968. Constraints on variables in syntax: MIT
dissertation. [Published in 1986 as Infinite Syntax! Norwood, N.
J.: Ablex].
Russell, Bertrand. 1905. On denoting. Mind 14. 479–93.
Strawson, P. F. 1950. On referring. Mind 59(235). 320–344.
Szabolcsi, Anna. 1987. Bound variables in syntax: Are there any?
In Proceedings of the sixth amsterdam colloquium, University of
Amsterdam.
190
BIBLIOGRAPHY
Vikner, Carl & Per Anker Jensen. 2002. A semantic analysis of the
English genitive: Interaction of lexical and formal semantics.
Studia Linguistica 56. 191–226.
Winter, Yoad. 2001. Flexibility principles in Boolean semantics.
Cambridge, MA: MIT Press.